to the monograph

Transcription

to the monograph
A Contribution to the Finite
Element Analysis of High-Speed
Compressible Flows and
Aerodynamic Shape Optimization
M. Kouhi
E. Oñate
G. Bugeda
Monograph CIMNE Nº-140,February 2013
A Contribution to the Finite
Element Analysis of High-Speed
Compressible Flows and
Aerodynamic Shape Optimization
M. Kouhi
E. Oñate
G. Bugeda
Monograph CIMNE Nº-140, February 2013
International Center for Numerical Methods in Engineering
Gran Capitán s/n, 08034 Barcelona, Spain
INTERNATIONAL CENTER FOR NUMERICAL METHODS IN ENGINEERING
Edificio C1, Campus Norte UPC
Gran Capitán s/n
08034 Barcelona, Spain
www.cimne.com
First edition: February 2013
A CONTRIBUTION TO THE FINITE ELEMENT ANALYSIS OF HIGH-SPEED COMPRESSIBLE FLOWS
AND AERODYNAMIC SHAPE OPTIMIZATION
Monograph CIMNE M140
 Los autores
ISBN: 978-84-941686-2-8
Depósito legal: B-14079-2013
Summary
This work covers a contribution to two most interesting research fields in aerodynamics, the finite element analysis of high-speed compressible flows (Part I)
and aerodynamic shape optimization (Part II).
The first part of this study aims at the development of a new stabilization
formulation based on the Finite Increment Calculus (FIC) scheme for the Euler
and Navier-Stokes equations in the context of the Galerkin finite element method
(FEM). The FIC method is based on expressing the balance of fluxes in a spacetime domain of finite size. It is tried to prevent the creation of instabilities
normally presented in the numerical solutions due to the high convective term
and sharp gradients.
In order to overcome the typical instabilities happening in the numerical
solution of the high-speed compressible flows, two stabilization terms, called
streamline term and transverse term, are added through the FIC formulation in
space-time domain to the original conservative equations of mass, momentum
and energy. Generally, the streamline term holding the direction of the velocity
is responsible for stabilizing the spurious solutions produced from the convective
term while the transverse term smooths the solution in the high gradient zones.
An explicit fourth order Runge-Kutta scheme is implemented to advance the
solution in time.
In order to investigate the capability of the proposed formulation, some
numerical test examples corresponding to subsonic, transonic and supersonic
regimes for inviscid and viscous flows are presented. The behavior of the proposed stabilization technique in providing appropriate solutions has been studied especially near the zones where the solution has some complexities such
as shock waves, boundary layer, stagnation point, etc. Although the derived
methodology delivers precise results with a nearly coarse mesh, the mesh refinement technique is coupled in the solution to create a suitable mesh particularly
in the high gradient zones.
The comparison of the numerical results obtained from the FIC formulation
with the reference ones demonstrates the robustness of the proposed method for
stabilization of the Euler and Navier-Stokes equations. It is observed that the
usual oscillations occur in the Galerkin FEM, especially near the high gradient
zones, are cured by implementing the proposed stabilization terms. Furthermore, allowing the adaptation framework to modify the mesh, the quality of
the results improves significantly.
The second part of this work proposes a procedure for aerodynamic shape optimization combining Genetic Algorithm (GA) and mesh refinement technique.
In particular, it is investigated the effect of mesh refinement on the computational cost and solution accuracy during the process of aerodynamic shape
optimization. Therefore, an adaptive remeshing technique is joined to the CFD
solver for the analysis of each design candidate to guarantee the production of
more realistic solutions during the optimum design process in the presence of
shock waves.
In this study, some practical transonic airfoil design problems using adap1
tive mesh techniques coupled to Multi-Objective Genetic Algorithms (MOGAs)
and Euler flow analyzer are addressed. The methodology is implemented to
solve three practical design problems; the first test case considers a reconstruction design optimization that minimizes the pressure error between a predefined
pressure curve and candidate pressure distribution. The second test considers
the total drag minimization by designing airfoil shape operating at transonic
speeds. For the final test case, a multi-objective design optimization is conducted to maximize both the lift to drag ratio (L/D) and lift coefficient (Cl). The
solutions obtained with and without adaptive mesh refinement are compared in
terms of solution accuracy and computational cost. These design problems under transonic speeds need to be solved with a fine mesh, particularly near the
object, to capture the shock waves that will cost high computational time and
require solution accuracy.
By comparison of the the numerical results obtained with both optimization
problems, the obtainment of direct benefits in the reduction of the total computational cost through a better convergence to the final solution is evaluated.
Indeed, the improvement of the solution quality when an adaptive remeshing
technique is coupled with the optimum design strategy can be judged.
2
Resumen
El presente trabajo pretende contribuir a dos de los campos de investigación
más interesantes en la aerodinámica, el análisis numérico de flujos compresibles
a alta velocidad (Parte I) y la optimización de la forma aerodinámica (Parte II).
La primera parte de este estudio se centra en la solución numérica de las
ecuaciones de Navier-Stokes, que modelan el comportamiento de flujos compresibles a alta velocidad. La discretización espacial se lleva a cabo mediante el
método de elementos finitos (FEM) y se pone especial énfasis en el desarrollo
de una nueva formulación estabilizada basada en la técnica de cálculo de Incremento finitos (FIC). En este última, los términos de estabilización convectiva se
obtienen de manera natural de las ecuaciones de gobierno a través de postulados
de conservación y equilibrio de flujos en un dominio espacio-tiempo de tamaño
finito. Ello lleva a la obtención de dos términos de estabilización que funcionan
de manera complementaria. Uno actúa en dirección de las lı́neas de corriente
proporcionando la estabilización necesaria para contrarestrar las inestabilidades
propias de la forma discreta de Galerkin y el otro término, de tipo shock capturing, actúa de manera transversal a las lı́neas de corriente y permite mejorar
la solución numérica alrededor de discontinuidades y otro tipos de fenómenos
localizados en el campo de solución de problema. La forma discreta de las ecuaciones de gobierno se completa mediante un esquema de integración temporal
explı́cito de tipo de Runge-Kutta de 4to orden. El esquema de solución básico
propuesto se complementa con una técnica de refinamiento adaptativo de malla
que permite mejorar automáticamente la solución numérica en zonas localizadas
del dominio en que, dadas las caracterı́sticas del flujo, se necesita una mayor
resolución espacial.
Con el propósito de investigar el comportamiento de la formulación numérica
se estudian diferentes casos de análisis que implican flujos viscosos y no viscosos
en régimen subsónico, transónico y supersónico y se estudia con especial detalle el funcionamiento de la técnica de estabilización propuesta. Los resultados
obtenidos demuestran una exactitud satisfactoria y una buena correlación con
resultados presentes en la literatura, incluso cuando se trabaja con discretizaciones espaciales relativamente gruesas. Adicionalmente, los estudios numéricos
realizados demuestran que el empleo del esquema adaptativo de malla es eficaz
para incrementar la exactitud de la solución numérica manteniendo un bajo
coste computacional.
En la segunda parte de este estudio se propone un método para la optimización de formas aerodinámicas que combina algoritmos genéticos multiobjetivo (MOGAs) y remallado adaptativo con el objetivo de asegurar, con un
coste computacional mı́nimo, la calidad de la solución numérica empleada en el
proceso de búsqueda de un determinado diseño objetivo, particularmente cuando
el flujo presenta discontinuidades y gradientes muy localizados, tı́picos de flujos
a alta velocidad. La metodologı́a se aplica a resolver tres problemas prácticos
de diseño de perfiles aerodinámicos en flujo transónico que implican la optimización de la distribución de presiones, minimización de la resistencia de onda y
maximización conjunta de la sustentación y la relación sustentación/resistencia.
3
Para cada uno de ellos se estudia el efecto del refinamiento en la calidad de la
solución numérica ası́ como también en el coste computacional y la convergencia
del problema. Los estudios realizados demuestran la eficacia de la metodologı́a
propuesta.
4
Acknowledgment
I would like to express my sincere and heartfelt gratitude to all my teachers
whose dedication and hard work have paved my academic path. I owe special
thanks to my advisors, Professor Eugenio Oñate and Professor Gabriel Bugeda,
whose patience and guidance were instrumental in kindling the flame of my
curiosity and whose encouragements gave me the courage to challenge my own
limitations.
During my Ph.D., I was blessed with the kindly help of Dr. DongSeop Lee
and Dr. Jordi Pons in the aerodynamic design part and Dr. Roberto Flores
and Enrique Ortega in the high-speed flow part. I learned a lot from them and
never forget their help.
I want to appreciate Dr. Pooyan Dadvand and Dr. Riccardo Rossi who
provided me the possibility of writing my codes in KRATOS and consistently
solved my programming problems. It was a grate chance for me to get familiar
with different aspects of the serial and parallel computing based on the C++
and Python languages.
I must also thank my family who bore the difficulty of separation for the
last five years and whose love and support gave purpose to all my endeavors.
A special thanks to all my friends, without whom life at Barcelona would be
utterly miserable.
Last but not least, I would like to thank again Professor Eugenio Oñate, as
the director of CIMNE, for giving the chance of doing my Ph.D. in such a high
level center and for the scholarship that funded my work during the last 5 years.
Your support enables plenty of interesting research to be carried out by young
students and represents a firm example of faith in the next generations.
5
Contents
1 Introduction
1.1 Motivation . . . . . . . . . . . . . . . . . . . . . . . . . .
1.2 Stabilization Methods for High-Speed Compressible Flows
1.3 Aerodynamic Shape Optimization . . . . . . . . . . . . .
1.3.1 Numerical Optimization Methodologies . . . . . .
1.4 Adaptive Mesh Refinement . . . . . . . . . . . . . . . . .
1.5 Overview of Work . . . . . . . . . . . . . . . . . . . . . .
1.5.1 Objectives . . . . . . . . . . . . . . . . . . . . . . .
1.5.2 Structure . . . . . . . . . . . . . . . . . . . . . . .
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
13
13
17
19
20
22
23
23
25
2 A Stabilized FEM for High-Speed Compressible Flows
2.1 Compressible Euler/Navier-Stokes Equations . . . . . . .
2.2 A FIC-based Stabilization Formulation . . . . . . . . . . .
2.2.1 FIC in Space . . . . . . . . . . . . . . . . . . . . .
2.2.2 FIC in Space-Time . . . . . . . . . . . . . . . . . .
2.2.3 The General FIC-based Formulation . . . . . . . .
2.3 Space-Time Discretization . . . . . . . . . . . . . . . . . .
2.3.1 Galerkin FE . . . . . . . . . . . . . . . . . . . . . .
2.3.2 The Fourth Order Runge-Kutta . . . . . . . . . . .
2.4 Boundary Conditions . . . . . . . . . . . . . . . . . . . . .
2.4.1 Euler Equation . . . . . . . . . . . . . . . . . . . .
2.4.2 Navier-Stokes Equation . . . . . . . . . . . . . . .
2.5 Performance Enhancement . . . . . . . . . . . . . . . . . .
2.5.1 Local Time Stepping . . . . . . . . . . . . . . . . .
2.5.2 Residual Smoothing . . . . . . . . . . . . . . . . .
2.6 Mesh Refinement . . . . . . . . . . . . . . . . . . . . . . .
2.6.1 h-refinement . . . . . . . . . . . . . . . . . . . . .
2.6.2 Adaptive Remeshing . . . . . . . . . . . . . . . . .
2.7 Test Examples . . . . . . . . . . . . . . . . . . . . . . . .
2.7.1 Inviscid Flow . . . . . . . . . . . . . . . . . . . . .
2.7.2 Viscous Flow . . . . . . . . . . . . . . . . . . . . .
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
26
26
29
29
30
32
32
32
34
36
36
36
37
37
37
38
38
39
40
41
57
6
3 Aerodynamic Shape Optimization Using Genetic Algorithm And
Adaptive Remeshing
69
3.1 Overall Design Process . . . . . . . . . . . . . . . . . . . . . . . . 69
3.2 Flow Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70
3.2.1 CFD Solver Enhanced by Adaptive Refinement . . . . . . 70
3.2.2 Validation of the mesh refinement technique . . . . . . . . 72
3.3 Optimization Methodology . . . . . . . . . . . . . . . . . . . . . 78
3.3.1 Multi-Objective Genetic Algorithm . . . . . . . . . . . . 78
3.3.2 Parametrization . . . . . . . . . . . . . . . . . . . . . . . 80
3.4 Realistic Optimization Test Cases . . . . . . . . . . . . . . . . . . 81
3.4.1 Reconstruction Design . . . . . . . . . . . . . . . . . . . . 82
3.4.2 Drag Minimization . . . . . . . . . . . . . . . . . . . . . . 88
3.4.3 Multi-Objective Design . . . . . . . . . . . . . . . . . . . 94
4 Concluding Remarks and Future Work
7
98
List of Figures
2.1
2.2
2.3
2.4
2.5
2.6
2.7
2.8
2.9
2.10
2.11
2.12
2.13
2.14
Reflected shock example. Problem definition. . . . . . . . . . . .
Reflected shock. (a) Initial mesh and (b) adaptive mesh after 5
refinement levels. . . . . . . . . . . . . . . . . . . . . . . . . . .
Reflected shock. The results using initial mesh (a) density contours and (b) pressure contours. . . . . . . . . . . . . . . . . . .
Reflected shock. The results using adaptive mesh (a) density
contours and (b) pressure contours. . . . . . . . . . . . . . . . .
Reflected shock example. Comparison of the density profiles at
y = 0.25. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Reflected shock. The results obtained from β = 0.25 (a) adaptive mesh after 5 refinement levels, (b) density contours using
uniform mesh, (c) density contours using adaptive mesh and (d)
comparison of the density profiles at y = 0.25. . . . . . . . . . .
Reflected shock. The results obtained from β = 0.75 (a) adaptive mesh after 5 refinement levels, (b) density contours using
uniform mesh, (c) density contours using adaptive mesh and (d)
comparison of the density profiles at y = 0.25. . . . . . . . . . .
Subsonic inviscid flow around around a NACA0012 airfoil example. (a) Domain discretization and (b) airfoil close-up. . . . . . .
Subsonic inviscid flow around around a NACA0012 airfoil example. (a) Density contours and (b) close-up density lines in the
stagnation area. . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Subsonic inviscid flow around around a NACA0012 airfoil example. (a) Convergence of the density at the stagnation point and
(b) density value along the stagnation streamline. . . . . . . . . .
Subsonic inviscid flow around around a NACA0012 airfoil example. Pressure coefficient contours. . . . . . . . . . . . . . . . . .
Transonic inviscid flow around a NACA0012 airfoil. Final adapted
mesh. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Transonic inviscid flow around a NACA0012 airfoil. Obtained
solution for the initial mesh. (a) density contours and (b) pressure
contours. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Transonic inviscid flow around a NACA0012 airfoil. Obtained
solution for the adaptive mesh. (a) density contours and (b)
pressure contours. . . . . . . . . . . . . . . . . . . . . . . . . . .
8
42
42
43
43
44
45
46
47
48
49
50
50
51
52
2.15 Transonic inviscid flow around a NACA0012 airfoil. The comparison of the cp distributions with the reference values. . . . . .
2.16 Supersonic inviscid flow around a NACA0012 airfoil. The adaptive mesh after (a) one level, (b) three levels and (c) five levels of
refinement. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.17 Supersonic inviscid flow around a NACA0012 airfoil. Obtained
solution for the initial mesh. (a) density and (b) mach number
contours. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.18 Supersonic inviscid flow around a NACA0012 airfoil. Obtained
solution for the refined mesh. (a) density and (b) mach number
contours. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.19 Supersonic inviscid flow around NACA0012 airfoil. The comparison of the cp distributions with the reference values. . . . . . . .
2.20 Subsonic laminar flow past NACA0012 airfoil. Detail of the mesh.
2.21 Subsonic laminar flow past a NACA0012 airfoil. Mach number
contours. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.22 Subsonic laminar flow past a NACA0012 airfoil. (a) Close-up of
computed velocity vectors near the trailing edge and (b) details
of pressure contours. . . . . . . . . . . . . . . . . . . . . . . . . .
2.23 Subsonic laminar flow past a NACA0012 airfoil. Comparison of
the obtained pressure coefficient Cp distribution with the numerical results of reference [18]. . . . . . . . . . . . . . . . . . . . . .
2.24 Subsonic laminar flow past a NACA0012 airfoil. Comparison
of the obtained skin-friction coefficient Cf distribution with the
numerical results of reference [18]. . . . . . . . . . . . . . . . . .
2.25 Subsonic laminar flow past a NACA0012 airfoil. Convergence of
the density at the stagnation point for different values of β . . .
2.26 Supersonic flow over flat plate. Problem definition. . . . . . . . .
2.27 Supersonic flow over flat plate. (a) Density, (b) pressure, (c)
temperature and (d) Mach number contours. . . . . . . . . . . .
2.28 Supersonic flow over flat plate. Comparison of the obtained (a)
density (b) vertical velocity profiles along the line x = 1.2 with
the reference results [11]. . . . . . . . . . . . . . . . . . . . . . . .
2.29 Compression corner. Problem definition. . . . . . . . . . . . . . .
2.30 Compression corner. Detail of the structured mesh. . . . . . . . .
2.31 Compression corner. (a) density, (b) pressure, (c) temperature
and (d) Mach number contours. . . . . . . . . . . . . . . . . . . .
2.32 Compression corner. Comparison of the obtained pressure coefficient Cp distribution with the numerical results of reference
[11]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.33 Compression corner. Comparison of the obtained skin-friction
coefficient Cf distribution with the numerical results of reference
[11]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.1
Flowchart of the design process. . . . . . . . . . . . . . . . . . . .
9
53
54
55
56
57
58
58
59
60
60
61
62
63
64
65
66
67
68
68
71
3.2
3.3
3.4
3.5
3.6
3.7
3.8
3.9
3.10
3.11
3.12
3.13
3.14
3.15
3.16
3.17
3.18
3.19
3.20
3.21
3.22
3.23
3.24
Initial mesh containing 2084 nodes and 3970 3-noded triangle
elements. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
The obtained mesh after (a) 2 (b) 4 (c) 6 (d) 10 (e) 20 (f) 30
remeshing levels for M∞ = 0.8 and α = 0.0◦ . . . . . . . . . . . .
Cp field after 20 refinement levels for M∞ = 0.8 and α = 0.0◦ . . .
Comparison between the Cp distribution after 30 remeshing levels
and numerical results reported in [45] for M∞ = 0.8 and α = 0.0◦ .
Total number of nodes and elements versus the refinement level
number for M∞ = 0.8 and α = 0.0◦ . . . . . . . . . . . . . . . . .
The obtained mesh after (a) 2 (b) 4 (c) 6 (d) 10 (e) 20 (f) 30
remeshing levels for M∞ = 0.85 and α = 1.0◦ . . . . . . . . . . . .
Cp field after 20 refinement levels for M∞ = 0.85 and α = 1.0◦ . .
Comparison between the Cp distribution after 30 refinement levels and numerical results reported in [49] for M∞ = 0.85 and
α = 1.0◦ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Total number of nodes and elements versus the refinement level
number for M∞ = 0.85 and α = 1.0◦ . . . . . . . . . . . . . . . . .
Topology of Robust Multi-objective Optimization Platform. . . .
Upper and lower bound of design variables for NACA 0012 (top)
and RAE 2822 (bottom) in comparison to the original ones and
corresponding free and fixed control points. . . . . . . . . . . . .
The baseline mesh around NACA 0012 5(a) and RAE 2822 5(b).
contours around NACA 0012 at the flow conditions M∞ = 0.78
and α = 2.0◦ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Convergence history of reconstruction design of uniform and adaptive mesh test cases. . . . . . . . . . . . . . . . . . . . . . . . . .
Comparison between target, uniform mesh and adaptive mesh
test case airfoils. . . . . . . . . . . . . . . . . . . . . . . . . . . .
Comparison between target, uniform mesh test case and adaptive
mesh test case Cp distributions. . . . . . . . . . . . . . . . . . . .
Cp contours around target airfoil (a), uniform mesh test case (b)
and adaptive mesh test case (c) in Cp range of [−1.21 : 1.19]. . .
Mach number contours around target airfoil (a), uniform mesh
test case (b) and adaptive mesh test case (c) in Mach range of
[0.0 : 1.38]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Convergence history for drag minimization for the uniform and
adaptive mesh test cases. . . . . . . . . . . . . . . . . . . . . . .
The comparison between the baseline RAE 2822, the adaptive
mesh and uniform mesh test case. . . . . . . . . . . . . . . . . . .
Cp distribution obtained from baseline, adaptive mesh and uniform mesh test cases. . . . . . . . . . . . . . . . . . . . . . . . . .
Cp contours around original airfoil (a), uniform mesh test case
(b) and adaptive mesh test case (c) in Cp range of [−1.45 : 1.17].
Mach number contours around original airfoil (a), uniform mesh
test case (b) and adaptive mesh test case (c) in Mach range of
[0.0 : 1.43]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
10
72
73
74
74
75
76
76
77
77
79
81
82
83
84
85
86
87
87
89
91
92
93
93
3.25 Optimized Pareto fronts after 150 generations and baseline design.
3.26 Geometry comparison between the baseline NACA 0012 and optimal airfoils. . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.27 Cp distribution obtained from baseline design and compromised
solution. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.28 Cp contours obtained by the baseline design (a) and compromised
solution (b) at Cp range of [−1.37 : 1.20]. . . . . . . . . . . . . .
11
95
96
97
97
List of Tables
2.1
3.1
3.2
3.3
3.4
3.5
3.6
3.7
Subsonic laminar flow past a NACA0012 airfoil. Comparison of
separation location values obtained from different values of β. . .
Comparison of fitness function values obtaining for adaptive mesh
and uniform mesh test cases. . . . . . . . . . . . . . . . . . . . .
Comparison of fitness function values obtaining for adaptive mesh
and uniform mesh cases eliminating the effect of mesh difference.
Comparison of fitness function values obtained from baseline RAE
2822, adaptive mesh and uniform mesh test cases. . . . . . . . .
Comparison of fitness function values obtained from baseline RAE
2822, adaptive mesh and uniform mesh cases eliminating the effect of mesh difference. . . . . . . . . . . . . . . . . . . . . . . . .
Airfoil configuration of the baseline RAE 2822 and optimized
airfoils using adaptive mesh and uniform mesh. . . . . . . . . . .
The comparison of the aerodynamic coefficients for adaptive remeshing approach and the baseline design NACA 0012. . . . . . . . .
Airfoil configuration of the baseline NACA 0012 and the optimized airfoils using adaptive remeshing approach. . . . . . . . . .
12
61
84
85
89
90
91
95
96
Chapter 1
Introduction
1.1
Motivation
Advances in computational methods over the past two decades have firmly motivated researchers to implement these schemes as a powerful tool to handle the
existing problems in different engineering areas. Aerodynamics is one of the
most important branch of engineering which always has been considered as a
highly interesting application topic for computational methods.
Applications of aerodynamics are common in the design of aircraft in aeronautical engineering, the design of race cars and railway trains in the automotive engineering, the study of wind effect on slender bridges and tall buildings in
structural engineering and many other engineering problems. Regarding the different aspects appearing in this area, a huge amount of investigation fields have
been created. The implementation of the numerical methods in two important
areas, high-speed compressible flows (Part I) and aerodynamic shape optimization (Part II), is investigated in this work, separately, and new methodologies
are developed for each one. A detailed introduction to each specific part follows.
Part I: High-Speed Compressible Flow
The theory of high speed flow is concerned with flows of fluid at speeds high
enough that the fluid0 s compressibility must be taken into account. The theory
finds application in many branches of science and technology such as aerodynamics of aircrafts and vehicles. In general, the physical behavior of compressible flows is more complicated than in incompressible flows. Compressible flows
may be viscous or inviscid depending on the flow viscosity. Compressible inviscid
flows are analyzed using the potential or Euler equations, whereas compressible
viscous flows are solved via the Navier-Stokes system of equations. Air flows in
the range (0.3 ≤ M ≤ 5.0) may be considered as compressible. This range is
usually subdivided into regions identified as subsonic (0.3 ≤ M ≤ 0.8), transonic
(0.8 ≤ M ≤ 1.2), and supersonic (1.2 ≤ M ≤ 5.0).
13
In high speed flow, the variables in the mass, momentum and energy equations are coupled to the thermodynamic variables, because changes in pressure
compress or expand the fluid and alter its temperature. Equally, changes in
temperature affect the pressure, via the equation of state. Therefore, in the
study of high speed flow there is no escape from some thermodynamics.
The theory of shock waves is an important part of the subject of high speed
flow and occupies an appreciable proportion of the researches carried out in
this field. Furthermore, shock wave turbulent boundary layer interactions in
compressible viscous flows constitute one of the most important physical phenomena in computational fluid dynamics. The occurrence of these complexities
have made a wide range of investigations in the classical levels of experimental,
theoretical and numerical computation.
An intense research effort into the technology of high speed flow has been
performed by some brilliant mathematicians, including Lighthill, Von Neumann
and Prandtl. A feature of research work on high speed flow has been the increasing use of high speed computers, hand-in-hand with the creation of a new
subject, computational fluid dynamics (CFD). Among many successes has been
the numerical computation of transonic flow fields, as required for the design of
transonic airfoils. The use of high speed computers now pervades all aspects of
research into high speed flow and, indeed, other types of flow.
The classical numerical methods such as central finite difference (FD), finite
volume (FV) and Galerkin finite element (FE) are the most popular methods
in the area of high-speed compressible flows. It has been studied that these
discretization methods suffers from the occurrence of the spurious solutions in
the numerical results due to the the presence of convective terms [38]. To prevent
the occurrence of these non-physical solutions, some stabilization terms must be
added to the basic scheme. Also, in order to model and capture the complexities
of the high-speed compressible flows (specially near the high gradient zones
such as shock waves), extra stabilization terms, called shock capturing terms, is
needed in the numerical simulation.
As the stabilization terms play an important role in the quality of the results obtained from the numerical methods, a huge amount of research has been
done in this area for predicting an appropriate stabilization term. In order to
address this challenge, the Finite Increment Calculus (FIC) method is employed
in conjunction with Galerkin FE as the basis for the derivation of stabilization
terms in this work. This method is capable of generating accurate results for
both inviscid and viscous high-speed compressible flows. In addition, the FIC
formulation has a sound mathematical foundation that differs from the ad hoc
procedures for introducing the stabilization term in many alternative stabilization procedures.
The FIC method is based in invoking the balance of fluxes in a fluid domain of
finite size. This introduces naturally additional terms in the classical differential
equations of infinitesimal fluid mechanics which are a function of the balance
domain dimensions. The new terms in the modified governing equations provide
the necessary stabilization to the discrete equations obtained via the standard
Galerkin finite element method. The implementation of the FIC procedure
14
in the stabilized form of the advective-diffusive transport and incompressible
fluid flow problems is investigated by Oñate and co-workers [80, 84, 85, 86, 87,
81, 83, 82]. The obtained results exhibit the robustness and efficiency of the
FIC formulation in the mentioned problems. Hence, the extension of the FIC
procedure to derive stabilized finite element schemes for high-speed compressible
flow problems is investigated in this work.
Besides an appropriate stabilization technique, a suitable computational
mesh can enhance magnifically the quality and the precision of the numerical results, specially around the zones where the gradient of the solution is high.
This fact motivated us to develop powerful technologies, generally called mesh
refinement, to improve the mesh quality around these zones. When using mesh
refinement methodologies, a minimum level of error is obtained by paying a reasonable amount of computational cost using a number of degrees of freedom as
small as possible. This is achieved by placing much more elements in the zones
where the solution has a greater gradient value or discretization error. Also, the
elements are removed from the regions where the solution has a smaller gradient value or discretization error to have a minimum computational cost. In this
work, the mesh refinement technique developed by Löhner [63] is implemented
in conjunction with the FIC formulation in order to obtain a better enrichment
level.
Part II: Aerodynamic Shape Optimization
Aerodynamic design is one the most important challenges in the engineering
problems. The main goal of aerodynamic design is to efficiently and accurately
determine an aircraft configuration that attains the optimal aerodynamic performance. The most significant parts of each configuration are the wings and
gas-turbine engines which are commonly subjected to be optimized.
According to the complexity of aerodynamic design, several methodologies
are developed in this field by scientists. Experimental methods, using wind
tunnel and flight testing, provide the basis of the traditional “cut and try” approach for the design of new aerodynamic shapes. This approach alone, however,
is too expensive, which has motivated the continuous development of sophisticated computational methods for flow simulation [66] and established the field of
computational fluid dynamics (CFD). These methods, along with performance
gains in computer technology, complement and in some instances even replace
the use of the wind tunnel during the design process. In this setting, the experimental and computational techniques are analysis tools that provide reliable
estimates of aerodynamic performance for given configurations and operating
conditions. A good physical insight of the designer is required to select and
evolve the candidate aerodynamic shapes and provide an overall control for the
design process.
The “cut and try” approach does not result in an inefficient design process.
Consequently, a significant research effort has been devoted to the development
of computational methods for the solution of the design problem. This leads to
the creation of “aerodynamic shape optimization” concept which is the coupling
15
of CFD with numerical optimization. The methods developed based on this
concept are efficient at producing aircraft shape configurations with improved
performance characteristics at a given aircraft operating condition.
In the early years, the lack of computational resources and robust algorithms
limited the overall impact of CFD on the design process. For a while, challenges
in accurately predicting fluid dynamic phenomena within reasonable amounts of
time were considered to be the foremost shortcoming hindering the widespread
use of numerical simulations.
The primary challenge of CFD in the aerodynamic shape optimization is
the underlying complex nature of the fluid flow. For cruise configurations of
commercial aircraft, the flow is compressible, usually transonic, and may contain features such as shock waves, shock-induced boundary-layer separation and
boundary-layer transition which imply using a fine mesh around the zones where
these complexities occur. Hence, mesh refinement seems to be interesting for
the aerodynamic shape optimization of some real test cases where the control
of the mesh is extremely significant. Hence, by utilizing mesh refinement, a
suitable mesh is generated around these zones to predict these complexities of
the flow in an accurate manner keeping the computational cost reasonable.
Although the most conventional approach of aerodynamic design problems is
to use a fixed uniform mesh during the optimization iterations [116, 96, 88, 40],
a few researches have been performed on implementing adaptive mesh for each
design candidate of the optimization. In this category, the works done by Bugeda
and Oñate [6, 7] can be mentioned where they developed a methodology which
utilizes adaptive mesh for each design in a suitable manner for a full potential
flow and Euler equation without shock waves. Their methodology is based on
the derivation of the sensitivity of the nodal coordinates and some flow variables
with respect to the design variables to project the remeshing parameters from
the old design to the new one.
The goal of this research is the integration of adaptive mesh refinement
technique with aerodynamic shape optimization under transonic flow regime in
order to show the benefits of using mesh adaptation in terms of computational
cost and solution accuracy when the solution of the fluid has the previously
mentioned complexities.
The rest of this chapter is organized as follows: In Section 1.2 a summary
on stabilization methodologies developed in the field of high-speed compressible
flows is presented. An overview of algorithms for aerodynamic shape optimization problems is given in Section 1.3 containing different aspects required in
a practical problem. Section 1.4 describes the concept of mesh refinement in
the engineering problems. The objectives of this work are stated in Section 1.5
which also provides an outline of the document.
16
1.2
Stabilization Methods for High-Speed Compressible Flows
In the past decades, computational resources and algorithms have matured to
a state such that numerical modeling is an essential component of engineering
analysis and design. This is certainly true for computational fluid dynamics
(CFD), which has grown into the ability to solve flow fields with sophisticated
geometries and complex physical processes. While experimental measurements
will always have a role in the design process, CFD offers advantages in terms
of cost, test time, ease of use and quality of output data. Nevertheless, despite
the advances in the usage and capabilities of CFD, there is still room for improvement. One area of CFD growth is in the development of accurate schemes
and their application to an expanding diversity of flow regimes and problems.
Because of the large amount of demands from the aerospace industry, the application of CFD schemes in the regime of high-speed compressible flows always
has been interested for the investigators.
As mentioned before, the main difficulties regarding the numerical methods
in high-speed compressible flows is the occurrence of numerical instability which
has two main sources, the high value of convective term in the original partial
differential equation and the sharp gradients in the solution. Much effort has
been spent in developing the so called stabilized numerical schemes for all the
standard numerical methods such as FD, FV and FE. The following is a brief
overview with an historical perspective of two main categories of the stabilization methods, artificial diffusion and limiters.
Artificial diffusion
To overcome the instabilities, the key idea is to introduce more diffusion in
the flow equations by adding viscous terms to the governing partial differential equations explicitly. This concept formed the base of a well known family
of stabilization methods called artificial diffusion technique proposed by Von
Neumann and Richtmyer in 1950 [79].
Within this context, the finite difference method was investigated by Courant
et al. in 1952 [17] and Lax in 1954 [56] to solve the high-speed compressible
flow problems numerically. Courant et al. [17] introduced the upwind family
of finite difference methods which was continued by Godunov in 1959 [33] for
developing a new finite difference method based on the solution of the so-called
Riemann problem. This original approach generated a series of schemes, known
as flux difference splitting methods, that introduce different approximate Riemann solvers proposed by Engquist and Osher [26], Roe [103, 102] and Osher
[89].
Lax [56] implemented the traditional first order finite difference as the numerical methods for discretization of the Euler and Navier-Stokes equations
whereas the development of the second order finite difference methods was provided by Lax and Wendroff [55] and Mac-Cormack [69] implementing an explicit
17
time integration while Lerat [57] presented the implicit one.
Based on the finite volume scheme, following the idea of artificial diffusion,
an important numerical improvement was conducted by Jameson et al. [52]
using a series of second and fourth order stabilization methods. The study
of finite volume flux vector splitting for the Euler equations was presented by
Anderson et al. [2] where several advantages of the MUSCL-type approach over
standard flux-differencing one is discussed.
The Galerkin finite element method is an alternative numerical procedure
which has been widely applied for the Euler and Navier-Stokes equations. Hughes
and his group [42] developed the classical Streamline-Upwind/Petrove-Galerkin
(SUPG) finite element method, initially proposed by Brooks and Hughes [5] for
incompressible flows, for high-speed compressible flows in the context of conservation variables. After that, in order to prevent the instabilities around the
discontinuities, they utilized a set of entropy variables in conjunction with a
new shock capturing operator depending on the residual of the corresponding
equation [41].
Beau and Tezduyar [32] have shown that by considering the same shock
capturing operator, the results obtained from [41] can be recovered by entropy
variables with the similar level of accuracy. Following the idea of the conservative variable formulation, Mittal and Tezduyar [77] presented a unified finite
element method for incompressible and compressible flows where a parameter
z, based on the local Mach number, was introduced that governs the choice of
equations for compressible and incompressible flows.
Based on the concept of the SUPG method, several stabilization techniques
were introduced such as Taylor-Galerkin [25] and Galerkin least squares (GLS)
methods [12, 108] which coincided with the original SUPG method under some
specified conditions. Also, the implementation of the GLS method with different
sets of variables can be found in the work of Hauke and Hughes [36] where they
mentioned that the choice of entropy variables leads to the most robust results
in the presence of singularities.
Peraire et al. [94] and Morgan et al. [78] developed a new artificial diffusion scheme based on the mesh size, h, and second order gradient of pressure.
Zienkiewicz and Wu [125] derived a new formulation by adding a pressure switch
obtained from the nodal pressure values inside the element. Using of fractional
step method [13, 14], Zienkiewicz and co-workers introduced a general fluid mechanics algorithm, called characteristic-based split (CBS) algorithm, for incompressible and compressible flows [123, 124] which benefited from the anisotropic
shock capturing term presented by Codina [15]. The semi-implicit form of the
CBS algorithm for compressible and incompressible flow is investigated by Codina and co-workers [16].
Limiters
Besides the artificial diffusion methods, a modern family of stabilization methods, based on the so-called limiters, was derived and has been commonly used
in finite volume and finite difference schemes. This method is based on a con18
cept aimed at preventing the generation of new extrema in the flow solution in
such a way that the values of local minima do not decrease, and the values of
local maxima do not increase [112, 113, 58]. This method induced to the Flux
Corrected Transport (FCT) scheme, developed by Boris and Book [4], and the
Total Variation Diminishing (TVD) method introduced by Harten [34].
A fully multidimensional generalization of the FCT algorithm was proposed
by Zalesak [122] and carried over to the finite element method by Parrott [92],
Löhner et al. [65] and Luo et al. [67].
Sanderos [105] developed the original TVD method from the explicit/implicit
fully discrete scheme to a semi-discrete one, whereas Jameson and Lax [50] derived a general TVD characterization and the necessary conditions for multipoint support in explicit, implicit and semi-discrete formulations. Although the
FCT scheme has not got a significant popularity, the TVD method has been
used mostly in finite difference [23, 119, 120], finite volume [1] and finite element
[68, 54] methods.
1.3
Aerodynamic Shape Optimization
Aerodynamic shape optimization methods can be divided into two categories,
namely, inverse design methods and numerical optimization methods. Inverse
design methods, first introduced in 1945 by Lighthill [60], are an established
approach for the determination of an airfoil shape that attains a given pressure
distribution. For example, an experienced designer is able to specify a liftpreserving pressure distribution that is shock-free for transonic flow conditions,
thereby achieving significant drag reductions. Inverse methods were also used
to design the well-known Liebeck high-lift airfoils [59]. An advantage of this
approach is its low computational cost. Giles and Drela [31] developed an
inverse design method using the two-dimensional, coupled Euler and boundary
layer equations with a computational cost equivalent to the solution of just
one analysis problem. Although inverse design methods have been applied to
three-dimensional problems [28], their primary limitation is the specification
of desirable pressure distributions that lead to optimal designs, especially for
problems with multiple operating conditions and turbulent and separated flow.
Numerical optimization methods provide a more general approach for solving design problems. The creativity and insight of an experienced designer are
required to reduce the design problem to a well-posed optimization problem.
This involves the definition of objective functions that specify the goals of the
optimization, design variables that determine the aerodynamic shape, as well
as constraints that qualify a feasible region of the design space. Note that for
practical problems, it is very likely that the objectives are competing and that
changes in the specification of the optimization problem occur as the design
evolves. Typically, the problem is cast as a minimization one, where the objective functions include lift, drag and moment functionals. Inverse design can
be considered a subset of numerical optimization by defining an objective function that represents the difference between the target and the actual pressure
19
distributions.
Once an aerodynamic shape optimization problem is defined , the following
steps provide the aerodynamic optimization process. A flow solver is used to
determine the properties of the flow field around an aerodynamic shape such as
lift, drag and pressure coefficient at a given set of operating conditions. Then,
the formulated objective function evaluates the performance of the shape with
respect to the design objectives. A mathematical representation of the geometry
of the shape provides a mean to make alterations to the shape via design variables. An optimization algorithm uses information about the objective function
at the current design iteration to determine how to modify the design variables to improve the performance of the shape. The updated shape specified
by the modified design variables is presented to the flow solver and the process is repeated iteratively until the specified criteria are satisfied indicating
that an optimal solution has been achieved and no further improvement in the
performance is possible.
1.3.1
Numerical Optimization Methodologies
Despite the fact that numerical optimization methods have been successfully
used for a countless number of design problems, an application of numerical
optimization to aerodynamic design still remains as a formidable challenge because of the following two difficulties: 1) objective function landscape of an
aerodynamic optimization is often multi-modal and nonlinear because the flow
field is governed by a system of nonlinear partial differential equations. and 2)
function evaluations using a CFD code, especially a three-dimensional Euler or
Navier-Stokes code, are very expensive. Due to the above difficulties, aerodynamic design problems require a numerical optimization tool to be very robust
and efficient as well.
Typical optimization problems involve the determination of the minimum of
a given function. The goal is to identify the values of the parameters that are
inputs for the function such that the output of the function is a minimum. If the
slope (or gradient) of the function at any point is known, it becomes possible to
move toward the minimum by moving in the direction of negative slope. The
gradient can also be interpreted as the sensitivity of the functional output to
the parameters that control the function. In the context of aerodynamic shape
design, functional quantities of interest are usually surface integrated values
such as the lift, drag, or moment. Design parameters usually are the variables
that control the shape of the geometry and the governing flow equations form
the function relating the design variables to the output functional. If gradientbased optimization is to be employed for the purpose of reducing the output
functional, the goal then is to determine the gradient or sensitivity of the output
with respect to the design variables.
The obvious choice for determining such sensitivities is to perturb the design variables individually and run the numerical simulation (evaluation of the
function) for each perturbation in order to determine the effect of the perturbation on the quantity of interest. The method, known as finite-differencing,
20
becomes inefficient as the number of design variables increases since the number
of times the simulation has to be performed is directly dependent on the number
of design variables. The sensitivity or the gradient determined in this manner
is also highly dependent on the magnitude of the perturbation, since for large
perturbations the wrong slope may be predicted due to nonlinear effects, while
very small perturbations are susceptible to machine precision related issues. In
a perfect world using a computer with infinite precision, an infinitesimally small
perturbation would recover the analytically exact value of the gradient. Hicks
et al. [37] used the finite-difference method to evaluate sensitivities.
Using control theory, the gradient can be calculated indirectly by solving an
adjoint equation. This methodology significantly lowers the computational cost
and is clearly an improvement over classical finite-difference methods. Although
there is the additional overhead of solving the adjoint equation, once it has been
solved the cost of obtaining the derivatives of the cost function with respect to
each design variable is negligible. Consequently, the total cost to obtain these
gradients is independent of the number of design variables and amounts to the
cost of one flow solution and one adjoint solution. The adjoint problem is
a linear partial differential equation of lower complexity than the flow solver.
Jameson was the first to apply control theory for transonic design problems
[46, 47, 48]. Subsequently, Jameson et al. [51] pioneered the shape optimization
method for Euler and Navier-Stokes problems. Automatic aerodynamic design
of aircraft configurations has yielded optimized solutions of wing and wing-body
configurations by Reuther et al. [100, 101], Burgreen et al. [9] and Löhner et
al. [99, 110].
Evolutionary Algorithms (EAs) are alternative optimization algorithms mimicking mechanism of the natural evolution, where a biological population evolves
over generations to adapt to an environment by selection, recombination and
mutation. When EAs are applied to optimization problems, fitness, individual
and genes usually correspond to an objective function value, a design candidate,
and design variables, respectively. One of the key features of EAs is that they
search from multiple points in the design space, instead of moving from a single
point like gradient-based methods do. Furthermore, these methods work on
function evaluations alone and do not require derivatives or gradients of the objective function. These methods also offer the benefit of being able to navigate
through highly nonlinear, noisy and discontinuous design environments. However, such methods usually require very large numbers of function evaluations
and present their own set of difficulties. These features lead to the advantages
such as robustness, suitability to parallel computing and simplicity in coupling
CFD codes. Owing to these advantages over the analytical methods, EAs have
become increasingly popular in a broad class of design problems (for example,
see [76]). EAs have been also successfully applied to aerodynamic shape optimization problems such as airfoil shape design [98, 118], Multi-element airfoil
shape design [10], subsonic wing shape design [88] and supersonic wing shape
design [90].
21
1.4
Adaptive Mesh Refinement
Adaptive methods for the solution of the governing flow equations are means
of reducing the overall cost of simulations. The basic principle behind adaptive
solutions is the efficient deployment of computational resources.
In the case of finite volume or finite element discretizations of the governing
flow equations, the computational expense of obtaining a solution is directly
proportional to the resolution of the mesh. However, for two different meshes
that have the same number of points, it is possible to obtain different error
levels in the overall solution due to different distributions of the mesh points.
Non-adaptive solutions rely on user judgment for the clustering of mesh points
in various regions of the computational domain in order to capture flow phenomena. The normal practice consists of using some predetermined clustering
of points in high gradient regions such as boundary layers, wakes, stagnation
points, shock waves and contact discontinuities. Since the flow field being solved
for is unknown, the user has to rely on engineering judgment for the determination of regions where clustering may be required. Adaptive solvers approach
this problem by starting with a coarse resolution mesh and then add points
on-the-fly during the solution process in relevant regions of the computational
domain.
The core of any adaptation algorithm is the determination of relevant regions
to target for increased resolution. A common adaptation method is to use the
spatial gradients of the evolving flow solution to refine regions of high gradients
in the hopes that the overall solution error level is reduced. If mesh adaptation
is implemented during the process of optimization, while the approach may
capture interesting flow phenomena such as vortices and shock waves, there
is no guarantee that the solution obtained in this manner necessarily results
in improved estimates for specific objective values of interest such as lift or
drag. To this end, goal-based or adjoint-based methods for the adaptation of
the computational domain have been developed to specifically target the error
in objective scalars of interest [8, 70]. Goal-based methods use a posteriori
estimates of the error in a functional computed using a flow solution to identify
regions in the computational domain that are most relevant to the accuracy of
the functional. For different scalars of interest computed using the same flow
solution (i.e. identical flow conditions and geometry), it is entirely possible that
goal-based adaptation produces significantly different adaptations.
In each adaption problem two main features must be considered; (i) A reliable error indicator, and (ii) A refinement methodology, where two features are
explained in the following.
Error indicator
In most adaptivity problems, the aim is to reach a uniform level of error in the
domain. This error can be evaluated in different manners. A major difficulty in
achieving definite improvements using adaptation for Euler and Navier-Stokes
calculations is the lack of a reliable error indicator [63, 73]. A common strategy
22
is to adapt to certain physical features of the flow, such as shock waves, boundary layers, wakes, slip lines, or stagnation points, by employing error indicators.
Normally, the error indicators depend on the gradient of a solution variable like
the Mach number, density or entropy. It is typical to classify the error indicators based on the order of the gradient. The indicators obtained from the
first order gradients performs well when the problem considered is inviscid in
nature [91, 74, 75]. Also, one can approximate the error in the domain using a
derivative one order higher than the shape function of the solution where the
derivatives are obtained by some recovery procedures [106].
Refinement methodology
Some powerful refinement strategies are introduced in the literature to properly refine the mesh. Four main directions are followed:
1. p-refinement that adds further degrees of freedom with hierarchical shape
functions or with the addition of higher order shape functions[29, 117]. This
method is more commonly used in variational frameworks such as DiscontinuousGalerkin discretizations, which use a high-order polynomial representation of the
solution within each element.
2. r-refinement (or mesh movement) that maintains the mesh topology but
allows mesh lines to move, thus contracting some cells while simultaneously expanding others. This technique has the advantage that the number of mesh
cells and hence computational cost does not increase when a new flow field is
calculated on the adapted mesh [104].
3. h-refinement (mesh enrichment or subdivision) where individual elements are
subdivided without altering their original position [114, 115, 73].
4. Adaptive remeshing that regenerates a new mesh by applying new element
sizes obtained from error indicators. A grid generator must be utilized in order
to regenerate the mesh [75, 61].
1.5
1.5.1
Overview of Work
Objectives
As mentioned at the beginning of this chapter, two important research fields in
aerodynamics, high-speed compressible flows (Part I) and aerodynamic shape
optimization (Part II), are considered as the main contributions of this work.
An overview of the methodologies proposed in each part of the work is presented
in the following lines:
Part I: High-Speed Compressible Flow
The first part of this research is dedicated to the development of the FIC formulation for stabilization of the Euler and Navier-Stokes equations in the context
23
of the Galerkin FEM. The FIC method is based on expressing the balance of
fluxes in the momentum, mass balance and energy conservation equations in
a space-time domain of finite size. It is intended to prevent the creation of
instabilities usually presented in the numerical solutions due to the high convective terms and sharp gradients. To reach this aim, two stabilization terms,
called the streamline term and the transverse term, are added through the FIC
formulation.
Generally, the streamline term which is in the direction of the velocity is
responsible for stabilizing the spurious solutions produced from the convective
terms while the transverse term smooths the solution in the high gradient zones
inside the domain. A fourth order Runge-Kutta scheme has been implemented
to advance the solution in time.
In order to investigate the capability of the proposed FIC-FEM formulation,
some numerical test examples corresponding to subsonic, transonic and supersonic regimes for inviscid and viscous flows are presented. The behavior of the
stabilization terms in providing appropriate solutions has been studied especially near the zones where the solution has some complexities such as shock
waves, boundary layer, stagnation point, etc. Although the derived methodology delivers precise results with a nearly coarse mesh, the mesh refinement
technique is coupled with the solution to create a suitable unstructured mesh
particularly near the high gradient zones.
The comparison of the numerical results obtained from the FIC-FEM formulation with the reference ones demonstrates the robustness of the proposed
method for stabilization of the Euler and Navier-Stokes equations for compressible flows. It is observed that the usual oscillations observed in the Galerkin
FEM, especially near the high gradient zones, are eliminated by implementing
the proposed stabilization terms without introducing an excessive numerical dissipation. Furthermore, allowing the adaptation framework to modify the mesh,
the quality of the results improves significantly.
Part II: Aerodynamic Shape Optimization
The second part of this work investigates the effect of mesh refinement on the
computational cost and solution accuracy during the process of aerodynamic
shape optimization. Therefore, an adaptive remeshing technique is linked to
the CFD solver for the analysis of each design candidate to guarantee the production of more realistic solutions during the optimum design process in the
presence of shock waves.
In this study, some practical transonic airfoil design problems using adaptive
mesh techniques coupled to Multi-Objective Genetic Algorithms (MOGAs) and
an Euler flow analysis code are addressed. The methodology is implemented to
solve three practical design problems; the first test case considers a reconstruction design optimization that minimizes the pressure error between a predefined
pressure curve and a candidate pressure distribution. The second test considers
the total drag minimization by designing the airfoil shape operating at transonic speeds. For the final test case, a multi-objective design optimization is
24
conducted to maximize both the lift to drag ratio (L/D) and the lift coefficient
(Cl). The mentioned design problems under transonic speeds need to be solved
with a fine mesh, particularly near the object, to capture the shock waves that
will cost high computational time and require solution accuracy.
The solutions obtained with and without adaptive mesh refinement are compared in terms of solution accuracy and computational cost. By comparison of
the numerical results, the direct benefits in the reduction of the total computational cost through a better convergence to the final solution are evaluated.
Indeed, the improvement of the solution quality when an adaptive remeshing
technique is coupled to the optimum design strategy is evident.
1.5.2
Structure
The structure of the work is as followings: Chapter 2 presents the derivation
of the stabilized Galerkin FEM based on the FIC formulation for analysis of
compressible flows. Next, the mesh refinement technique is introduced and some
numerical results for inviscid and viscous compressible flows are shown at the
end of the chapter. Chapter 3 addresses the aerodynamic shape optimization
part of the work. First, the optimization tools such as the GA technique and the
parametrization method are presented. Then, the validation of the CFD solver
chosen for the solution of the compressible flow equations in conjunction with
the presented mesh refinement technique is demonstrated and finally a number
of optimum shape design examples are shown. Conclusions and areas of future
work are summarized in Chapter 4.
25
Chapter 2
A Stabilized FEM for
High-Speed Compressible
Flows
In this chapter, the FIC formulation is applied to finite element discretization of the Euler/Navier-Stokes equations for compressible flows. The proposed
method is demonstrated on a series of two-dimensional subsonic, transonic and
supersonic test cases involving various configurations. A commonly used error
indicator [63] in conjunction with the h-refinement methodology is implemented
to comparatively assess the performance of the proposed stabilized formulation.
In all the delivered test cases, the FIC scheme avoids the creation of instabilities which are inherently inside the solution due to the high convective term
and sharp gradients.
The first part of this chapter, Section 2.1, describes the original equations of
the Euler and Navier-Stokes. Section 2.2 presents the development of the FIC
stabilized formulation for the mentioned equations. The discretization of the
general formulation in space and time is delivered in Section 2.3. The different
kind of boundary conditions considered in this work are studied in Section 2.4.
Additional techniques for enhancing the performance of the method via local
time stepping and residual smoothing are presented in Section 2.5 while the
mesh refinement methodologies are described in Section 2.6. The numerical
test examples containing inviscid and viscous flows are shown in Section 2.7.
2.1
Compressible Euler/Navier-Stokes Equations
The Euler/Navier-Stokes equations are a system of partial differential equations
that describe the behavior of a compressible, inviscid/viscous fluid. They are
derived from the integral form of the laws of conservation of mass, momentum
and energy in an inertial frame of reference.
26
For an arbitrary fixed control volume Ω with the boundary Γ, the law of
conservation of mass can be expressed as
Z
Z
∂
ρ dΩ = − ρ(vj nj ) dΓ
(2.1)
∂t Ω
Γ
where ρ is the density, vj is the fluid velocity, expressed using indicial notation
and nj is the outward-pointing unit normal vector at the surface of the control
volume. This states that the rate of change of the mass of the fluid in the control
volume is equal to the transport of mass across the control volume boundary Γ.
Gauss’ divergence theorem is used to transform the surface integral into a volume integral over Ω. Then, by requiring the resulting integral equation to hold
for any infinitesimal control volume, the differential form of mass conservation
law is found to be
∂
∂ρ
+
(ρvj ) = 0
∂t
∂xj
(2.2)
which holds everywhere that the flow quantities are continuous and differentiable. At discontinuities, only the integral form is valid.
The integral form of the law of conservation of momentum can be written
as
Z
Z
Z
Z
∂
ρvi dΩ = − ρvi (vj nj ) dΓ − pni dΓ + σij nj dΓ
(2.3)
∂t Ω
Γ
Γ
Γ
where p and σ are the static pressure and the viscous stress tensor of the
fluid respectively. The index i spans the two equations, and the repeated index
j indicates summation. These equations state that the momentum of the fluid
within the control volume is changed by the transport of momentum across the
surface, and by the action of fluid pressure on the surface Γ. Again, the divergence theorem is used to transform the surface integral terms. The differential
form of the momentum equation is
∂(ρvi )
∂
∂p
∂
+
(ρvi vj ) +
−
(σij ) = 0
∂t
∂xj
∂xi
∂xj
(2.4)
Again, the differential form is not valid at flow discontinuities.
The integral form of energy conservation law can be written as
Z
Z
Z
Z
∂T
∂
ρe dΩ = − (ρe + p)(vj nj ) dΓ + k(
nj ) dΓ + (σij vi nj ) dΓ (2.5)
∂t Ω
∂xj
Γ
Γ
Γ
with e standing for the total internal energy per unit mass, T for the absolute
static temperature and k for the thermal conductivity coefficient. This states
that the energy within the control volume is changed by the transport of energy
across the surface, and by the work done by the action of the fluid pressure
upon the surface of Ω. By application of the divergence theorem, the energy
equation can be transformed into its differential form,
∂
∂
∂T
∂(ρe)
+
([ρe + p]vj ) −
(σij vi + k
)=0
∂t
∂xj
∂xj
∂xj
27
(2.6)
which is only valid in continuous regions of the flow.
Since the three conservation laws have analogous terms, they can be grouped
together to form a system of equations,
∂Gi
∂Φ ∂Fi
+
−
= 0 for i = 1, 2
∂t
∂xi
∂xi
(2.7)
where Φ is the vector of conservative variables. By defining Ui = ρvi , this vector
has the following form


ρ
 U1 

(2.8)
Φ=
 U2 
ρe
with F and G being the vectors of inviscid and viscous flux respectively,





Ui

v
U
Fi = 
 i 1 + pδi1
 vi U2 + pδi2
vi (p + ρe)


Gi = 







0
σ1i
σ2i
∂T
k ∂x
+ vj σij
i





(2.9)
The Navier-Poisson law for a Newtonian fluid leads to the components of
the viscous stress tensor σ in term of the velocity. For an isotropic media
σij = µ(
∂vi
∂vj
∂vk
+
)+λ
δij for k = 1, 2
∂xj
∂xi
∂xk
(2.10)
where a bulk viscosity of (3λ + 2µ) = 0 is assumed, which is justified by the
validity of the Stokes hypowork for reversible processes under compression. µ =
µ(T ) is the dynamic coefficient of viscosity which is calculated from Sutherland‘s
equation.
The momentum and energy equations are coupled by the pressure which is
defined by assuming that the fluid behaves as a perfect gas. Hence, the pressure
is obtained from the equation of state for a perfect gas as
p = (γ − 1)ρ(e − 0.5vj vj )
(2.11)
where the ratio of specific heats γ is defined as
γ=
Cp
Cv
(2.12)
In the present computations γ is taken to be equal to 1.4 for the simulation of
the air, which is an adequate choice for subsonic, transonic and supersonic flow
where the Mach number is no excessively high and where chemical reactions can
be neglected.
It is necessary to mention that the Euler equations can be recovered from the
Navier-Stokes equations by neglecting the viscous stress and the heat conduction
terms. i.e. Gi = 0.
28
2.2
A FIC-based Stabilization Formulation
In this section the FIC formulation for stabilization of the Euler/Navier-Stokes
equations for compressible flows will be presented. The FIC formulation in
space is developed for the the momentum and energy conservation equations
whereas the representation of the FIC in space-time is suggested for the mass
conservation equation. By joining the three conservation equations, the general stabilized formulation of the Euler/Navier-Stokes equations will be finally
presented.
2.2.1
FIC in Space
We define rmi and re as the residuals of the momentum and energy equations,
respectively, as
rmi :=
re :=
∂
∂p
∂
∂Ui
+
(ρvi vj ) +
−
(σij ) = 0
∂t
∂xj
∂xi
∂xj
∂
∂(ρe)
∂
∂T
+
(vj (ρe + p)) −
(σij vi + k
)=0
∂t
∂xj
∂xj
∂xj
(2.13)
(2.14)
where i, j = 1, nd with nd is the number of space dimensions (nd = 2 for 2D
flow problems), the FIC formulation in space for stabilization of the momentum
and energy equations can be found by applying Taylor series expansions to the
conservation laws of momentum and energy over a space domain as [80, 82]
1
rmi − hmi ∇rmi = 0
(2.15)
2
1
(2.16)
re − he ∇re = 0
2
and he are characteristic length parameters that will be discussed
where hmi
later.
It can be seen that the modified equations via the FIC method introduces
naturally an additional term into the standard momentum and energy equations.
The definition of the the characteristic length arises from the two main sources
of the instabilities in the numerical solutions of high-speed compressible flows,
namely the high value of the convective terms and the sharp gradients. By
considering this fact, the general form of the characteristic length is expressed
as
h = hv + hg
v
(2.17)
where the streamline length h is responsible for smoothing the instabilities
due to the high convective terms by adding extra diffusion in the streamline
direction, whereas the transverse length hg stabilizes the oscillations near the
sharp gradient zones by applying extra diffusion in the transverse direction to
the gradient.
29
Definition of the streamline length hv . The basic idea behind the evaluation
of hv is extracted from the traditional SUPG scheme for stabilization of the
incompressible/compressible flow problems, where the diffusion is added in the
direction of the velocity. Hence, the characteristic length hv corresponding to
the momentum and energy equations can be represented as
v
|v| + vc
hv mi = βmi `mi
hv e = βe `e
v
|v| + vc
(2.18)
where βmi and βe are constant coefficients, `mi and `e are characteristic element sizes corresponding to the momentum
and energy equations. Also, |v|
q
is the module of the velocity and vc = γ ρp is the speed of the sound in the flow.
Definition of the streamline length hg . As the transverse term is supposed to
get involved in the zones where there are some high gradients in the solution,
hg can be defined as
hg mi = (1 − βmi )`mi
∇vi
|∇vi |
hg e = (1 − βe )`e
∇T
|∇T |
(2.19)
By considering Equation 2.17, the summation of Equations 2.18 and 2.19 gives
v
∇vi
+ (1 − βmi )`mi
|v| + vc
|∇vi |
(2.20)
v
∇T
+ (1 − βe )`e
|v| + vc
|∇T |
(2.21)
hmi = βmi `mi
he = βe `e
Substituting the characteristic lengths hmi and he in Equations 2.15 and
2.16, the FIC formulation in space for the momentum and energy equations is
obtained as
1
∇vi
1
v
rmi − (1 − βmi )`mi
∇rmi − βmi `mi
∇rmi = 0
2
|∇vi |
2
|v| + vc
1
∇T
1
v
re − (1 − βe )`e
∇re − βe `e
∇re = 0
2
|∇T |
2
|v| + vc
2.2.2
(2.22)
(2.23)
FIC in Space-Time
The residual of the mass conservation equation rd can be expressed as
rd :=
∂ρ ∂Ui
+
=0
∂t
∂xi
(2.24)
with i = 1, nd where nd is the number of space dimensions (nd = 2 for 2D flow
problems).
As mentioned before, the FIC formulation in space-time is written in order to
derive the stabilization terms corresponding to the mass equation. In the same
30
manner as for the derivation of the FIC equations in space domain, the spacetime formulation can be introduced by considering the mass balance equation in
a space-time domain. Hence, after relatively simple algebra, the FIC formulation
for mass balance equation can be expressed as [80, 82]
1
1 ∂rd
rd − hd ∇rd + τd
=0
(2.25)
2
2 ∂t
where hd and τd are the characteristic length vector and the time stabilization
parameter corresponding to the mass balance equation, respectively, which will
be defined later.
It can be seen that the space and time derivatives of the residual corresponding to the mass equation are coupled together in the stabilized equation. By
substituting rd from Equation 2.24 in the time-derivative part of Equation 2.26
and retaining the term related to the space derivatives, the following expression
can be obtained
1 ∂2ρ 1
∂ ∂Ui
1
=0
rd − hd ∇rd + τd 2 + τd
2
2 ∂t
2 ∂xi ∂t
(2.26)
2
In Equation 2.26, although the term ∂∂t2ρ corresponding to the second derivative
of the density respect to time can be obtained explicitly using a simple backward
finite difference scheme, it is found that this term does not play an important
role in the stabilized formulation and will be neglected here onwards.
i
Replacing the term ∂U
∂t from Equation 2.13 into Equation 2.26 gives
1
1
rd − hd ∇rd − τd ∇(∇.(Fm − Gm )) = 0
(2.27)
2
2
where ∇.(Fm − Gm ) is the the divergence of the flux term corresponding to the
momentum equation with the form of
∇.(Fm − Gm ) =
∂p
∂
∂(ρvi vj )
+
−
(σij )
∂xj
∂xi
∂xj
(2.28)
Considering the same idea as the one used for the momentum and energy
equations to define the characteristic lengths, the expressions of hd and τd for
the mass balance equation can be defined as
hd = (1 − βd )`d
∇ρ
|∇ρ|
τd = βd
`d
|v| + vc
(2.29)
where βd is a constant and `d is the characteristic element size related to the
mass balance equation.
Substituting Equation 2.29 into Equation 2.27, the FIC stabilized formulation for the mass equation is obtained as
1
∇ρ
1
1
rd − (1 − βd )`d
∇rd − βd `d
∇(∇.(Fm − Gm )) = 0
2
|∇ρ|
2
|v| + vc
31
(2.30)
2.2.3
The General FIC-based Formulation
Having introduced the stabilized formulation for the mass, momentum and energy equations, we present the general formulation for the Euler/Navier-Stokes
equations. In the current work, accurate results have been obtained by considering the assumptions
`d = `m1 = `m2 = `e = `
(2.31)
βd = βm 1 = βm 2 = βe = β
(2.32)
where ` is a characteristic element size defined as ` = (2Ωe )1/2 where Ωe is the
element area for 2D problems and β = 0.5.
By using Equations 2.31 and 2.32 and joining Equations 2.30, 2.22 and 2.23,
the general FIC-based stabilized formulation can be expressed as
Mass balance
∇ρ
1
1
1
∇rd − β`
∇(∇.(Fm − Gm )) = 0
rd − (1 − β)`
2
|∇ρ|
2 |v| + vc
(2.33)
Momentum
1
∇vi
1
v
rmi − (1 − β)`
∇rmi − β`
∇rmi = 0
2
|∇vi |
2 |v| + vc
(2.34)
∇T
1
v
1
∇re − β`
∇re = 0
re − (1 − β)`
2
|∇T |
2 |v| + vc
(2.35)
Energy
Equations 2.33, 2.34 and 2.35 are the starting point for deriving the discrete
form of the stabilized Euler/Navier-Stokes equations in space and time. Clearly,
for the infinitesimal case l = 0, the standard balance equations 2.24, 2.13 and
2.14 can be recovered from the general stabilized formulation. It is noticeable
that expressing the stabilization terms as a function of the residuals of the
corresponding balance equations enforces the consistency of the proposed FIC
method.
2.3
2.3.1
Space-Time Discretization
Galerkin FE
Now, we can introduce a standard finite element discretization of the conservative variables by choosing C 0 continuous linear interpolation over triangle
elements as
n
X
Φ ' Φ̄ =
NJ Φ̄J
(2.36)
J=1
where n = 3 for the case of triangle elements, Φ̄ is the vector containing the
approximation of the conservative variables Φ, N is the matrix of the linear
32
interpolating shape functions and (.)J denotes Jth nodal values inside the elements.
By taking W as the vector of weight functions and applying the standard
weighted residual method to Equations 2.33, 2.34 and 2.35 and integrating by
parts the stabilization terms, neglecting the boundary terms, the variational
form of the discretized equations is found as
Z
W.r̄dΩ +
Ω
nel Z
X
e
Ωe
nel Z
¯
X
τ
∂W
ν ∂W ∂ Φ̃
.
dΩ +
Ai
.r̄st dΩ = 0
2 xi xi
xi
Ωe 2
e
(2.37)
where nel is the number of the elements and i = 1, 2. In the following, each
part of Equation 2.37 will be defined. The residual vectors r̄ and r̄st presented
in Equation 2.37 are




 r̄d

r̄ = 
 r̄m1
 r̄m2
r̄e





r̄st
 ∇.(F̄m − Ḡm )

r̄m1
=


r̄m2
r̄e





(2.38)
where r̄d , r̄m1 , r̄m2 and r̄e denote the approximate finite element residuals for
the mass, momentum and energy equations, respectively, and ∇.(F̄m − Ḡm ) is
the divergence of the approximate finite element flux term corresponding to the
momentum equation.
¯ in Equation 2.37 is the vector of approximated primitive
Furthermore, Φ̃
variables of the form


 ρ̄ 


¯

Φ̃ = 
 v̄x 
 v̄y 
T̄
(2.39)
β`
is the the time stabilization parameter and the matrices ν and
Also, τ = |v̄|+v̄
c
Ai have the following form

|r̄1 |
˜ |
|∇Φ̄
1

 0

ν = (1 − β)` 
 0

0

1
0
Ai = 
0
0

0
0
0
|r̄2 |
˜ |
|∇Φ̄
2
0

0 


0 

0
v̄i
0
0
33
0
|r̄3 |
˜ |
|∇Φ̄
3
0
0
0
0
v̄i
0

0
0

0
v̄i
(2.40)
|r̄4 |
˜ |
|∇Φ̄
4
(2.41)
where |.| denotes the absolute value.
Based on the Galerkin approximation, the weighting functions are assumed
to be equal to the interpolation ones, W = N. Indeed, by applying integration
by parts on the first term of Equation 2.37 we obtain
Z
Z
Z
∂N
∂ Φ̄
dΩ =
.(F̄i − Ḡi )dΩ − N.(F̄n − Ḡn )dΓ
N.
∂t
Ω xi
Γ
Ω
(2.42)
+
nel Z
¯
X
τ
∂N
ν ∂N ∂ Φ̃
.
dΩ +
Ai
.r̄st dΩ
2 xi xi
2
xi
Ωe
e
nel Z
X
Ωe
e
with i = 1, 2 and F̄n and Ḡn are the approximated inviscid and viscous flux
terms applied on the boundaries. In Section 2.4, the different types of the
boundary conditions will be defined.
It can be seen that in the right hand side of Equation 2.42, the first integral
represents the Galerkin term in a weak form, the third integral corresponds
to the elemental contribution of the streamline stabilization term and the last
integral is the elemental contribution of the shock capturing stabilization term.
2.3.2
The Fourth Order Runge-Kutta
After discretization of the Euler/Navier-Stokes equation in space and assembling the element contributions from Equation 2.42, the found global system of
discretized equations can be written as
MIJ
∂ Φ̄nJ
= RnI
∂t
(2.43)
with I, J = 1, nnode where nnode is the total number of nodes inside the domain
and (.)n means the value computed in time step n. In the above equation, MIJ
is the consistent finite element mass matrix
Z
MIJ =
NI .NJ dΩ
(2.44)
Ω
Also, RnI is the contribution of the Ith global node from the right hand side of
Equation 2.42 in time step n which has the form
Z
n
Z
∂NI
RnI =
.(F̄i − Ḡi )dΩ − NI .(F̄n − Ḡn )dΓ
Ω xi
Γ
(2.45)
+
(n Z
el
X
e
Ωe
nel Z
¯
X
ν ∂NI ∂ Φ̃
τ
∂NI
.
dΩ +
Ai
.r̄st dΩ
2 xi xi
2
xi
Ω
e
e
)n
To avoid solving a linear system of equations at each time step, the consistent
mass matrix M is usually replaced by its lumped expression ML . Furthermore,
34
as the focus of this work is on stationary problems, an explicit multi-stage
Runge-Kutta algorithm is implemented in order to obtain a converged steady
state solution,
Assuming that the nodal values Φ̄nJ and RnJ are known at time tn , the
advance of the solution over the time step tn to tn+1 is according to
(0)
Φ̄J = Φ̄nJ
..
.
(k)
(k−1)
Φ̄J = Φ̄nJ + αk ∆t[ML ]−1 RI
..
.
k = 1, ..., K
(2.46)
(K)
Φ̄n+1
= Φ̄J
J
with K being the number of stages of the scheme. Each particular scheme is
characterized by a choice of K and the constant coefficients αk . The appropriate
choose of these coefficients improves the stability of the time integration and
provides accurate numerical solutions. Reasonable results can be obtained by
K = 4 =⇒ α1 = 1/4; α2 = 1/3; α3 = 1/2; α4 = 1
(2.47)
The scheme presented here is explicit and therefore only conditionally stable.
There is an upper limit on the allowable time step. For an inviscid problem, the
stability limit for an element could be calculated as [38]
∆te = C
h
|v̄| + v̄c
(2.48)
where C denotes the allowable Courant number and h = ` is the characteristic
element size. Except C which is global, the remaining variables in the above
equation are calculated at the elemental level. If a time accurate solution is
sought, the global time step equals to the minimum allowable one for all the
elements in the mesh. Including the viscous terms, the critical time step can be
represented as
h
∆te = C
(2.49)
4µγ 3/2 M∞
|v̄| + v̄c + ρ̄min
P rRe∞ h
with ρ̄min is the minimum density within the element, M∞ is the free stream
Mach number, P r is the non dimensional Prandtl number and Re∞ is the free
stream Reynold number. The other variables have the same meaning as defined
above. Details can be found in [38]. In this work, the Prandtl number is assumed
to be constant and equal to 0.72.
In order to enhance the performance of the presented time integration scheme,
the methods of local time stepping and residual smoothing are coupled to the
solver. A brief review of these schemes is presented in Section 2.5.
35
2.4
2.4.1
Boundary Conditions
Euler Equation
The discretized equation system of Equation 2.42 assumes a computational domain Ω surrounded by a boundary Γ with unit normal n. So far, the algorithm
only describes the contributions of each element across the integral Ω but does
not yet states how to incorporate the boundary conditions at the boundary Γ.
In this work, two types of boundaries exist: the slip boundary ΓW through
which mass flux is not possible and the far field (inflow/outflow) boundary Γ∞
through which mass flux is possible. The boundary condition must be applied
in a compatible form with the equations to be solved.
Slip Boundary
The normal component of the velocity must vanish on it. This condition can
be enforced by considering the normal component of the velocity equal to zero
after each stage, k, of the time integration scheme as
v(k) .n = 0 on ΓW
(2.50)
Far Field Boundary
Depending on the regime of the flow, the components of the solution which enter
the domain are to be enforced and the ones leaving the domain have to be set
free. By using Roe0 s approximation Riemann solvers, the appropriate boundary
flux for node I located at the far field boundary is computed as
F̄In =
1
{F̄n (Φ̄I ) + F̄n (Φ̄∞ ) − |Ān (Φ̄I , Φ̄∞ )|(Φ̄I − Φ̄∞ )}
2
(2.51)
where the superscript ∞ represent the freestream and Ān (Φ̄I , Φ̄∞ ) is the Roe
matrix computed in the direction normal to the boundary. More details about
the derivation of the Roe matrix can be found in [111, 38].
2.4.2
Navier-Stokes Equation
The treatment of the boundary condition for the Navier-Stokes equations is similar than for the Euler equations. However the steady momentum and energy
equations are elliptic and its modeling is more complex. Details are given in [38].
Far Field Boundary
Numerically, for a node I located at the far field boundary, the flux ḠIn for node
I belongs to the far field boundary, can be obtained by applying the far field
values at the boundary, i.e. ḠIn = Ḡ∞
n .
No Slip Boundary
In the case of the Navier-Stokes equations, in addition to the conditions on the
36
normal velocity, some conditions must be considered for the temperature. The
physical no slip boundary conditions for the velocity is
vi = 0
(2.52)
where i = 1, nd . For temperature boundary conditions, if an adiabatic wall is
modeled, then the heat flux qn across the wall is zero as
qn = −k
∂T
=0
∂n
(2.53)
whereas, if an isothermal wall is given, the temperature is set
T = TW
(2.54)
where TW is a specified wall temperature.
2.5
2.5.1
Performance Enhancement
Local Time Stepping
In order to enhance the performance of the program and to accelerate the solution towards steady state, local time stepping can be used. Hence, time steps of
different sizes are used within each element, individually, according to the local
stability limit. As the mesh size increases, ie. with the distance from the body,
the time increment will also increase. However, this option is only permissible
if the transient solution is not of interest.
The local time stepping is implemented at the nodal level. For ith node in
the mesh, a nodal size hinode is calculated as
hinode = min(hjel )
(2.55)
where hjel is the size of the jth element connected to the ith node. The above
equation states that the size of node i is the minimum size of all the elements
to which i belongs. The allowable step at each node is then calculated by
substituting Equation 2.55 in to Equations 2.48 and 2.49 while the rest of the
variables are calculated at the nodal level.
2.5.2
Residual Smoothing
To increase the allowable time step, the residual smoothing technique can be
implemented. A Laplacian smoothing is included to extend the support of the
interpolation functions, yielding an increase in the allowable Courant number.
Hence, the smoothed residual r̂i is defined as
X
r̂i = r̄i + (r̂j − r̂i ) for all j connected to i
(2.56)
j
37
where r̄i is the residual of the ith node defined in Equation 2.38 and is the
smoothing coefficient assuming to 0.1.
It can be seen that solving the Equation 2.56 exactly for the smoothed residual r̂i would require solving a linear system of equations negating the advantages
of the explicit scheme. However, the smoothing is only used as a mean to achieve
faster convergence and has no effect on the steady state solution. Therefore, a
rough approximation to the value of r̂i is enough to obtain the desired result.
The approximate value of the smoothed residual is computed by means of a
Jacobi iteration scheme as
P
r̄i + j r̂jn−1
i
P
(2.57)
r̂n =
1+ j1
where subscript n denotes the iteration step.
In common applications a good result can be obtained by running two passes
of the Jacobi iteration scheme. The residual smoothing does not need to be
applied at each stage of the Runge-Kutta scheme. For example, in the case of
the four-stage scheme, smoothing just in the first and third stages is usually
enough to produce a twofold increase in the allowable Courant number.
2.6
Mesh Refinement
A powerful error indicator introduced by Löhner [63] is implemented in this
research. For each variable of interest U, we define the error as
error(U ) =
h2 |second derivative of U |
h|f irst derivative of U | + cn |mean value of U |
(2.58)
By dividing the second derivatives of the variable of interest by the absolute
value of the first derivatives, the error indicator becomes bounded, dimensionless
and the ”eating up” effect of strong features is avoided. The term following cn is
added as a noise filter in order not to refine wiggles or ripples which may appear
due to loss of monotonicity. The error value thus depends on the algorithm
chosen to solve the PDEs describing the physical process at hand. The variable
U is one of the flow variables such as the density, the Mach number, the velocity
modulus, etc. In the current work, the density has been chosen.
In the following, a brief overview of two methodologies for mesh refinement
implemented in this research, h-refinement and adaptive remeshing, is presented.
2.6.1
h-refinement
By far the most successful mesh enrichment strategy has been h-refinement
[63, 62, 61]. In the present research, this strategy has been implemented in
combination with the error indicator mentioned above. The multidimensional
form of this error indicator is
s
P R
I
J
2
k,l ( Ω N,k N,l dΩ.UJ )
I
E (U ) = P R
(2.59)
I
J
J
2
k,l ( Ω |N,k |[|N,l UJ | + cn |N,l ||UJ |] dΩ)
38
where N I denotes the shape function of node I and k, l = 1, 2. The fact that
this error indicator is dimensionless allows the simultaneous of several indicator
variables. Because the error indicator is bounded 0 ≤ EI ≤ 1, it can be used for
whole classes of problems without having to be scaled to the problem at hand.
This results in an important increase in user-friendliness, allowing non- expert
users access to automatic self-adaptive procedures. This error indicator has
been used successfully for many years on a variety of applications [62, 61, 64].
The main variant used to date achieve mesh enrichment/coarsening through
is the classic subdivision of elements into four (2D) by dividing each edge of
element into two.
The first step in the h-refinement methodology is to identify the element
required refinement or coarsening. By subdividing or removing the elements
corresponding to the refinement or coarsening, respectively, the new mesh is
created. Finally, the flow variables are interpolated from the old mesh to the
new one.
2.6.2
Adaptive Remeshing
The third family of refinement strategies is based on the existence of automatic
grid generators. This strategy is based on the existence of automatic grid generators. The grid generator is used in combination with an error indicator based
on the present solution technique to remesh the computational domain, either
globally or locally, in order to produce a more suitable discretization.
In adaptive remeshing strategy, it is necessary to obtain a more precise
estimation of the required element size and shape. The non-dimensional error
indicator introduced in Equation 2.58 can also be generalized for this strategy.
By defining the following derivative tensors
Z
I
(D0 )Ikl = h2 cn
|N,k
||N,lJ ||UJ | dΩ
Ω
Z
1 I
2
I
(D )kl = h
|N,k
||N,lJ UJ | dΩ
(2.60)
Ω
Z
I
(D2 )Ikl = h2 |
N,k
N,lJ UJ dΩ|
Ω
The error indicator on the present (old) grid E old is given by
E old =
(D2 )Ikl
(D1 )Ikl + (D0 )Ikl
which yields an error matrix E of the form

Exx Exy
E =  Eyx Eyy
Ezx Ezy
39

Exz
Eyz 
Ezz
(2.61)
(2.62)
It is assumed that the new element size hnew is proportional to old element
size hold by a factor ζ which is defined as
ζ=
hnew
hold
(2.63)
The improved error related to the new mesh has the form shown in the following
equation, i.e.
ζ 2 (D2 )Ikl
(E new )Ikl =
(2.64)
ζ(D1 )Ikl + (D0 )Ikl
Given the desired error indicator value E new for the improved mesh, the mesh
reduction factor ζ is given by
q
old
new (D 1 )I +
[(D1 )Ikl ]2 + 4(D0 )Ikl EEnew [(D1 )Ikl + (D0 )Ikl ]
1E
kl
I
ζkl =
(2.65)
2 E old
[(D1 )Ikl + (D0 )Ikl ]
In 2D case, ζxx and ζyy are obtained for each element, the minimum of these
two values is replaced in Equation 2.63 to calculate the new element size hnew .
It is worth noting that in the current methodology only a new element size
is prescribed for each element and stretching is not considered. This value is
assigned to the corresponding element in the background mesh to generate a new
one. By predefining the minimum and maximum element sizes, the computed
element size is checked to be in this desirable range. If this condition is not
satisfied, the minimum or maximum element size is considered.
An automatic grid generator is needed to generate the new mesh using the
information obtained from the old mesh. Any of the automatic grid generation
techniques currently available (advancing front [95, 93], Delaunay triangulations
[3, 74] and modified quadtree/octree [121, 109]) may be employed to regenerate the mesh. The most robust one which is implemented here and uses the
advancing front technique. In this work, given the minimum element size and
the desired error, several remeshing steps are performed every predefined time
steps of the solution process in order to guarantee a fine mesh at the final step
of the analysis.
The first step of each remeshing level is to evaluate the indicator corresponding to each element. Utilizing the obtained indicators, the new element size is
calculated for each element. The new mesh in created by implementing the
advancing front technique considering the old mesh as the background mesh.
Finally, the flow variables are interpolated from the old mesh to the new one.
2.7
Test Examples
In order to assess the performance of the outlined stabilization methodology,
this section presents several numerical examples of compressible inviscid/viscous
flow in subsonic, transonic and supersonic regimes for steady-state problems. All
the tests are performed with triangular meshes utilizing an explicit Runge-Kutta
40
time integration scheme where the relative L2 norm of the density residual is
taken as a criterion to check convergence. It is to mention that the computations
start by using the upstream values as the initial solution and they are stopped
after a reduction of four order of magnitude in the relative L2 norm of the
density residual.
For the cases for which the analytical solution exists, the obtained numerical
results are compared with exact solutions whereas for the ones that there is
not an analytical solution, the comparison is carried out with some numerical
references.
2.7.1
Inviscid Flow
The computations of the inviscid test cases are performed through the entirely
unstructured mesh which is enhanced by the h-refinement technique, presented
in Section 2.6.1. It is to mention that the refinement is not implemented in the
subsonic example as no shock wave happens in the solution.
At the beginning stage of the refinement, the solution starts using an unstructured mesh. Having reached the stationary solution, consecutive refinement
levels are carried out every 200 time-steps. By obtaining the final adaptive mesh,
the solution stops when the stationary point is gotten. It is notable that the
initial mesh, considered for the inviscid examples, must be fine enough to be
able to capture the main characteristics of the flow in order to detect the regions
where the refinement is needed.
Example I: Reflected shock
A popular example for supersonic regime is the reflected shock problem involving an oblique shock at the angle of 29◦ and its reflection from the boundary.
The main feature of this example is that it can be solved analytically. Hence,
it is possible to test the accuracy of the numerical results.
The problem consists in an uniform flow with Mach 2.0 at the angle of 10◦
in a rectangle domain of height 4.1 and length 1.1 which involves three flow
regions, plotted in Figure 2.1 schematically, as
Region 1
ρ = 1.0
M = 2.9
v1 = 2.9
v2 = 0.0
vc = 1.0
p = 0.7143
Region 2
ρ = 1.7
M = 2.3781
v1 = 2.6193
v2 = −0.5063
vc = 1.1218
p = 1.5282
Region 3
ρ = 2.6872
M = 1.9423
v1 = 2.4015
v2 = 0.0
vc = 1.2363
p = 2.9340
On the left and upper side of the domain the flow variables of density, velocity
and temperature have fixed values corresponding to the Region 1 and Region 2,
respectively, whereas the lower wall is considered as a no slip boundary where
41
the normal component of the velocity is assigned zero. The flow variables on
the right side of the domain have been left free.
Mach=2.378
Region 2
Mach=2.9
Mach=1.942
Region 3
Region 1
Figure 2.1: Reflected shock example. Problem definition.
The initial mesh, shown in Figure 2.2a, is generated by using an unstructured
mesh consisting of 1376 nodes and 2580 3-noded triangular elements. The final
adaptive mesh of 3352 nodes and 6456 elements is obtained after five steps of
refinement as shown in Figure 2.2b. The form of this mesh clearly demonstrates
that the refinement has been carried out along the flow discontinuities.
(a)
(b)
Figure 2.2: Reflected shock. (a) Initial mesh and (b) adaptive mesh after 5
refinement levels.
Figures 2.3 and 2.4 display numerical results corresponding to the initial
and final mesh, respectively, which indicate the smoothness of the solution in
all over the domain especially near the shocks. It can be seen that although
the FIC method is capable to predict appropriate results by using a coarse
discretization, the refinement enhances the resolution of the shocks. An angle
42
of approximately 29◦ is obtained by using both discretizations which means
that the shock locations are captured accurately. It can be mentioned that the
shocks are captured within four or five elements for both discretizations.
(a)
(b)
Figure 2.3: Reflected shock. The results using initial mesh (a) density contours
and (b) pressure contours.
(a)
(b)
Figure 2.4: Reflected shock. The results using adaptive mesh (a) density contours and (b) pressure contours.
43
The comparison of the density profiles at y = 0.25 corresponding to the exact
solution and the numerical solution obtained using the initial and the adaptive
meshes is depicted in Figure 2.5. Good agreement with the exact solution is
obtained.
3.0
2.5
Density
2.0
1.5
Initial mesh
1.0
Adaptive mesh
Exact
0.5
0
1
2
3
4
x
Figure 2.5: Reflected shock example. Comparison of the density profiles at
y = 0.25.
The capability of the present method for different values of the constant
coefficient β is tested in this test case. The numerical results corresponding
to β = 0.25 and β = 0.75 are displayed in Figures 2.6 and 2.7, respectively,
containing the final adapted meshed, density contours using the uniform and
the adapted meshes and the comparison of the density profiles at y = 0.25. It
can be found that the choice of β = 0.25 results in a more diffusive solution
while a sharper shock is obtained by assuming β = 0.75. Although the shock
positions are almost captured by the both uniform and adapted meshes, some
oscillations are seen in the solutions near the shocks. These figures justify the
choice of β = 0.5.
44
(a)
(b)
(c)
3.0
2.5
Density
2.0
1.5
Adaptive mesh
1.0
Initial mesh
0.5
0
1
2
3
4
x
(d)
Figure 2.6: Reflected shock. The results obtained from β = 0.25 (a) adaptive
mesh after 5 refinement levels, (b) density contours using uniform mesh, (c)
density contours using adaptive mesh and (d) comparison of the density profiles
at y = 0.25.
45
(a)
(b)
(c)
3.0
2.5
Density
2.0
1.5
Adaptive mesh
1.0
Initial mesh
0.5
0
1
2
3
4
x
(d)
Figure 2.7: Reflected shock. The results obtained from β = 0.75 (a) adaptive
mesh after 5 refinement levels, (b) density contours using uniform mesh, (c)
density contours using adaptive mesh and (d) comparison of the density profiles
at y = 0.25.
46
Example II: Subsonic inviscid flow around a NACA0012 airfoil
This example, taken from [124], illustrates the quality of the flow solution for an
inviscid subsonic compressible flow past a NACA0012 airfoil at M∞ = 0.5 and
α = 0.0◦ . A circular computational domain is considered which is discretized
by a mesh of 4539 nodes and 8680 3-noded triangular elements as shown in
Figure 2.8. Since the flow does not involve any shock waves, mesh refinement
is not employed in this example. The slip boundary condition is applied on the
surface of the airfoil whereas the far field boundary condition is considered on
the outer boundary.
(a)
(b)
Figure 2.8: Subsonic inviscid flow around around a NACA0012 airfoil example.
(a) Domain discretization and (b) airfoil close-up.
In Figure 2.9 the density contours with a zoom at the stagnation point are
presented demonstrating that no oscillations are observed, neither in the global
solution nor at the stagnation point. A test for accuracy is the value of the
density at the stagnation point for which the analytical solution is known, as
ρ0 = ρ∞ (1 +
1
γ − 1 2 γ−1
M∞ )
2
(2.66)
where ρ∞ = 1 and γ = 1.4 for this example. Inserting M∞ = 0.5, we obtain
ρ = 1.1297 as the analytical value. The numerical result for this test case was
obtained as 1.1321 which differs about 2% from the analytical value.
47
(a)
(b)
Figure 2.9: Subsonic inviscid flow around around a NACA0012 airfoil example.
(a) Density contours and (b) close-up density lines in the stagnation area.
The convergence history of the density at the stagnation point is presented
in Figure 2.10a showing that the FIC formulation does not present spurious
oscillations in the density values at the stagnation point. The corresponding
value along the stagnation streamline is given in Figure 2.10b proving that a
smooth solution is obtained in this zone. The pressure coefficient obtained can
be observed in Figure 2.11.
48
Density at stagnation point
1.14
1.12
1.10
Computed value
1.08
1.06
1.04
1.02
1.00
0
500
1000
1500
2000
2500
Steps
(a)
1.20
1.15
Density
1.10
1.05
1.00
0.95
0.90
-7
-6
-5
-4
-3
-2
-1
0
Distance x
(b)
Figure 2.10: Subsonic inviscid flow around around a NACA0012 airfoil example.
(a) Convergence of the density at the stagnation point and (b) density value
along the stagnation streamline.
49
-0.4
-0.2
Cp
0.
0.2
0.4
0.6
0.8
1.
0.
0.2
0.4
0.6
0.8
1.
x
Figure 2.11: Subsonic inviscid flow around around a NACA0012 airfoil example.
Pressure coefficient contours.
Example III: Transonic inviscid flow around a NACA0012 airfoil
This example demonstrates the ability of the FIC-FEM formulation for the
analysis of the transonic compressible flow around a NACA0012 airfoil at M∞ =
0.8 and α = 1.25◦ . This example is taken from the reports of the AGARD
working group 07 [97]. The initial mesh utilized for this example is the same
as the one shown in Figure 2.8. The slip boundary condition is assigned on
the surface of the airfoil whereas the far field boundary condition is applied on
the outer boundary. The adapted mesh after five steps of refinement, containing
8010 nodes and 15768 elements, is shown in Figure 2.12 where the concentration
of the elements in the vicinity of the high gradient zones is clearly seen.
Figure 2.12: Transonic inviscid flow around a NACA0012 airfoil. Final adapted
mesh.
50
The solution variables corresponding to the initial mesh and the adaptive
mesh are presented in Figures 2.13 and 2.14 from which the effect of mesh
refinement in improving the quality of the results can be observed. It can be
found that, using both initial and adaptive discretizations, the stronger shock
at the upper side of the airfoil is captured with minor oscillations as well as the
weaker one at the lower side.
(a)
(b)
Figure 2.13: Transonic inviscid flow around a NACA0012 airfoil. Obtained
solution for the initial mesh. (a) density contours and (b) pressure contours.
51
(a)
(b)
Figure 2.14: Transonic inviscid flow around a NACA0012 airfoil. Obtained
solution for the adaptive mesh. (a) density contours and (b) pressure contours.
The pressure coefficient distribution over the airfoil resulting from the initial
and adaptive meshes is graphically compared with the results of [97] in Figure
2.15 proving that accurate results have been obtained. Although the position
of the shock does not change using the remeshing scheme, the adapted mesh
improves the solution in the high-gradient zones.
52
-1.
Cp
-0.5
0.
0.5
Initial mesh
Adaptive mesh
AGARD
1.
0.
0.2
0.4
0.6
0.8
1.
x
Figure 2.15: Transonic inviscid flow around a NACA0012 airfoil. The comparison of the cp distributions with the reference values.
Example IV: Supersonic inviscid flow around a NACA0012 airfoil
The examples involves the supersonic flow around a NACA0012 at M∞ = 1.2
and α = 0.0◦ which is again evoked from the AGARD working group 07 [97].
The domain, the initial mesh and the boundary conditions are the same as for
the Example II. Using a similar scheme for the mesh refinement, the final refined
mesh is presented in Figure 2.16 containing 12245 nodes and 24759 elements.
The obtained solution from initial and adapted mesh are presented in Figures
2.17 and 2.18, respectively. It can be found that although suitable results have
been obtained using the initial mesh, the refinement scheme improves the quality
of the results, especially near the shock waves. Also, both meshes are able to
capture the shock waves near the leading and trailing edge of the airfoil.
53
(a)
(b)
(c)
Figure 2.16: Supersonic inviscid flow around a NACA0012 airfoil. The adaptive
mesh after (a) one level, (b) three levels and (c) five levels of refinement.
54
(a)
(b)
Figure 2.17: Supersonic inviscid flow around a NACA0012 airfoil. Obtained
solution for the initial mesh. (a) density and (b) mach number contours.
55
(a)
(b)
Figure 2.18: Supersonic inviscid flow around a NACA0012 airfoil. Obtained
solution for the refined mesh. (a) density and (b) mach number contours.
56
The pressure coefficient distributions over a NACA0012 surface obtained for
this example are shown in Figure 2.19. A good agreement with the reference
results can be observed using both meshes.
-0.5
-0.25
0.
Cp
0.25
0.5
0.75
Initial mesh
1.
Adaptive mesh
AGARD
1.25
1.5
0.
0.2
0.4
0.6
0.8
1.
x
Figure 2.19: Supersonic inviscid flow around NACA0012 airfoil. The comparison
of the cp distributions with the reference values.
2.7.2
Viscous Flow
For the viscous examples, the computational mesh is generated by a combination
of unstructured and structured meshes in such a way that a structured mesh is
created a priori near the solid boundaries merging with an unstructured mesh
at the remaining parts of the domain. This approach has been widely used for
viscous flow problems [44, 35]. However, the method has not enough flexibility
for the inclusion of mesh refinement or for the extension to general complex 3D
configurations. It is to mention that if the domain problem in small enough,
the structured mesh is generated in all over the domain without embedding the
unstructured mesh.
Example V: Subsonic laminar flow past a NACA0012 airfoil
The subsonic viscous flow around a NACA0012 airfoil is presented here for
demonstrating the behavior of the developed stabilized formulation in viscous
regime. The flow conditions are Re = 5000, M∞ = 0.5 and α = 0◦ . The
assumed circular domain is discretized into 12623 nodes and 25300 3-noded triangle elements including a structured mesh of 15 layers near the airfoil boundary
which is merged with an unstructured mesh in the remaining of the computational domain. For the first layer of elements at the boundary, the normal
57
element size has the value of 0.0005 which is increased by a geometric progression for the following layers. The details of the mesh near the airfoil are shown
in Figure 2.20. The no slip adiabatic wall condition is imposed at the airfoil
surface, whereas the far field condition is applied at the outer boundary.
Figure 2.20: Subsonic laminar flow past NACA0012 airfoil. Detail of the mesh.
Figure 2.21: Subsonic laminar flow past a NACA0012 airfoil. Mach number
contours.
The obtained Mach number contour is presented in Figure 2.21 showing an
overall excellent agreement with the reference result [18]. Figure 2.22a illustrates the recirculation bubble at the trailing edge. Each vector of the figure
represents the modulus and direction of velocity at each node of the mesh. Pres58
sure contours are shown in Figure 2.22b. The fact that the lines are parallel to
each other with almost no oscillations indicates the good quality of the results.
(a)
(b)
Figure 2.22: Subsonic laminar flow past a NACA0012 airfoil. (a) Close-up of
computed velocity vectors near the trailing edge and (b) details of pressure
contours.
A more severe test of accuracy is the plot of the pressure coefficient cp and the
skin friction cf along the airfoil, presented in Figures 2.23 and 2.24, respectively,
showing the accuracy of the obtained results in comparison with the reference
ones [18]. It is to be noted that the peak value of cf is slightly underestimated.
Better results can be obtained by using a finer mesh near the leading edge.
59
-1.
-0.5
Cp
0.
0.5
Current method
Reference method
1.
0.
0.2
0.4
0.6
0.8
1.
x
Figure 2.23: Subsonic laminar flow past a NACA0012 airfoil. Comparison of
the obtained pressure coefficient Cp distribution with the numerical results of
reference [18].
0.20
Current method
0.15
Reference method
Cf
0.10
0.05
0.00
-0.05
-0.10
0.0
0.2
0.4
0.6
0.8
1.0
x
Figure 2.24: Subsonic laminar flow past a NACA0012 airfoil. Comparison of
the obtained skin-friction coefficient Cf distribution with the numerical results
of reference [18].
The variation of the accuracy and the convergence with the change in the
60
constant coefficient β is investigated in this example. Table 1 presents an estimate of the solution accuracy as measured by the computed values of the
separation location versus using three different values of 0.25 0.50 and 0.75 for
β. It can be found that the values obtained from β = 0.25 and β = 0.50 have a
good agreement with the results presented in the reference paper [18], ranging
from 80.9% − 83.4% chord.
Table 2.1: Subsonic laminar flow past a NACA0012 airfoil. Comparison of
separation location values obtained from different values of β.
Separation Location
β = 0.25
82.4%
β = 0.50
83.0%
β = 0.75
91.0%
The variation of the convergence history of the density at the stagnation
point with the change in β is presented in Figure 2.25 showing that the choice
of β = 0.5 does not present spurious oscillations in the density values at the
stagnation point.
Β = 0.25
Β = 0.50
Β = 0.75
1.24
Density at stagnation point
1.22
1.20
1.18
1.16
1.14
1.12
1.10
0
2000
4000
6000
8000
10 000
Steps
Figure 2.25: Subsonic laminar flow past a NACA0012 airfoil. Convergence of
the density at the stagnation point for different values of β
Example VI: Supersonic flow over flat plate
The Carter’s flat plate example with the flow conditions of Re = 1000, M∞ =
3.0 and α = 0◦ is selected here to examine the capability of the current method
in the presence of shock waves and boundary layers. A rectangular domain is
presented in Figure 2.26, with the dimensions of 1.4 and 0.8 along the x and y
directions, respectively, where the leading edge of the plate is located at x = 0.2
61
and y = 0.0. The Reynolds number is calculated based on the free stream values
and unique length. A structured mesh is created for the example by division of
the domain in 64 parts and 112 parts in x and y directions, respectively.
y
=1.0, v x=1.0,v y=0.0,T =2.769 E −4
M =3.0
R e=1000
Shock wave
=1.0
v x =1.0
v y =0.0
T =2.769 E−4
Boundary layer
v y=0.0
 xy=q y=0.0
v x =0.0, v y=0.0, T=7.754 E−4
x
Figure 2.26: Supersonic flow over flat plate. Problem definition.
As shown in Figure 2.26, all the characteristic variables of ρ, vx , vy and T
are fixed at the inflow and upper sides of the domain since these boundaries
behave as a supersonic inlet. The no slip boundary condition is applied on the
surface of the plate whereas the stagnation temperature of
Tstag = T∞ (1 +
γ−1 2
M∞ )
2
(2.67)
is imposed there, as well. Although a prescription of the density is needed at
the subsonic part of the outflow boundary, the flow variables are left free there.
The obtained density, pressure, temperature and Mach number contours
are plotted in Figure 2.27 proving the appropriate behavior of the presented
formulation in capturing the shock wave and boundary layer.
62
(a)
(b)
(c)
(d)
Figure 2.27: Supersonic flow over flat plate. (a) Density, (b) pressure, (c)
temperature and (d) Mach number contours.
63
In order to have a more precise test, the obtained density value and the y
component of the velocity profiles along the line x = 1.2 are compared in Figures 2.28a and 2.28b, respectively, with the ones presented in [11]. Although the
obtained peak point values of both the density and y component of the velocity
profiles are not coincident with the reference ones, a good agreement with the
reference results can be observed.
1.8
Current method
Reference method
1.6
Normalized density
1.4
1.2
1.0
0.8
0.6
0.4
0.0
0.2
0.4
0.6
0.8
y-coordinate
(a)
0.15
Current method
Reference method
Vertical velocity
0.10
0.05
0.00
0.0
0.2
0.4
0.6
0.8
y-coordinate
(b)
Figure 2.28: Supersonic flow over flat plate. Comparison of the obtained (a)
density (b) vertical velocity profiles along the line x = 1.2 with the reference
results [11].
64
Example VII: Compression corner
This example is another benchmark of the FIC-FEM formulation for supersonic
viscous regimes again extracted from the reference [11] where the flow with the
conditions of Re = 16800, M∞ = 3.0 and α = 0◦ enters the domain passing
over a flat plate and then a compression corner of 10◦ including the shock wave
and boundary layers initiated from the leading edge of the plate. The Reynolds
number is calculated based on the free stream values and the distance between
the leading edge of the plate and the compression corner.
As shown in Figure 2.29 schematically, the computational domain of 0.0 ≤
x ≤ 1.9 and 0.0 ≤ y ≤ 0.716 is assumed where the leading edge of the flat plate
is located at x = 0.1 and the compression corner starts from x = 1.1.
The flow variables of density, velocity and temperature are fixed at the inflow
and upper boundaries where no condition is prescribed on the outflow boundary.
On the plate surface, the no-slip boundary condition, as well as the specification
of the temperature as the stagnation temperature, calculated from Equation
2.67, are applied.
y
M =3.0
R e=16800
=1.0, v x =1.0, v y =0.0,T =2.769 E−4
Shock wave
=1.0
v x =1.0
v y =0.0
T =2.769 E−4
v y =0.0
xy =q y =0.0
Boundary layer
v x =v y =0.0, T =7.754 E−4
x
Figure 2.29: Compression corner. Problem definition.
The domain is discretized using a structured mesh of 3-noded triangles,
shown in Figure 2.30, containing 200 points in the streamline direction and 50
points in the vertical direction where the minimum element size above the plate
is taken as 0.0011 giving the maximum aspect ratio of almost 10.
65
Figure 2.30: Compression corner. Detail of the structured mesh.
The obtained density, pressure, temperature and Mach number contours are
presented in Figure 2.31. The figure shows that the methodology developed in
this work is able to provide smooth results in all over the domain especially
near the shock wave as well as near the boundary layers. The only inaccuracy
observed in the results is the presence of non-realistic values at the zone close
to the stagnation point which is a point of singularity. It can be seen that the
weakness of the formulation in determining the temperature at the stagnation
point results in an overestimation of the Mach number there. This problem can
be resolved by using elements with less aspect ratio around that region.
The comparison of the obtained Cp and Cf distributions along the plate
surface with the ones presented by Carter [11] is shown in Figure 2.32 and
Figure 2.33, respectively. Generally, a good agreement is observed except for
the peak values at the stagnation point, as mentioned before. The location of
the separation point happens at x = 0.89 showing an appropriate compatibility
with the results presented in the [11], [107] and [43] ranging from x = 0.84 to
x = 0.89.
66
(a)
(b)
(c)
(d)
Figure 2.31: Compression corner. (a) density, (b) pressure, (c) temperature and
(d) Mach number contours.
67
0.25
Current method
Reference method
0.20
Cp
0.15
0.10
0.05
0.00
0.0
0.5
1.0
1.5
2.0
x-distance from leading edge
Figure 2.32: Compression corner. Comparison of the obtained pressure coefficient Cp distribution with the numerical results of reference [11].
0.025
Current method
0.020
Reference method
Cf
0.015
0.010
0.005
0.000
-0.005
0.0
0.5
1.0
1.5
2.0
x-distance from leading edge
Figure 2.33: Compression corner. Comparison of the obtained skin-friction
coefficient Cf distribution with the numerical results of reference [11].
68
Chapter 3
Aerodynamic Shape
Optimization Using Genetic
Algorithm And Adaptive
Remeshing
The main objective of this chapter is to compare the solutions obtained from
different optimization methods with and without the use of adaptive remeshing
technique. The solutions will be compared in terms of accuracy and computational cost. Two different methods are considered. The first uses the MultiObjective Genetic Algorithm (MOGA) technique associated with adaptive mesh
refinement technique while the second uses the MOGA coupled with a conventional mesh technique (uniform mesh). Both use an Euler flow code [30] and a
multi-objective optimization software [19, 21, 20] developed at CIMNE in recent
years and they are implemented to three practical CFD design problems of the
airfoils in transonic regime. The main objective is to compare the solutions obtained with and without adaptive mesh refinement in terms of solution accuracy
and computational cost.
This chapter is organized as follows. Section 3.1 presents the overall design process. Section 3.2 describes the fluid solver and the mesh refinement
techniques which are coupled to it, including a validation test example. Section 3.3 presents the numerical optimization methods implemented in this work,
i.e., the MOGA technique and the parametrization methodology. A number of
optimization test cases are presented in Section 3.4.
3.1
Overall Design Process
A design code is developed in the current work which can be modularized into
several components such as the flow solver, the adaptive mesh refinement solver,
69
the parametrization algorithm and the optimization algorithm. The CIMNE
in-house codes named PUMI and RMOP are used as the flow solver and the
optimization algorithm, respectively, while the adaptive mesh refinement solver
and the parametrization algorithm are written in this work.
For the uniform mesh method, the design procedure can be described as
follows
1. Define initial design variables.
2. Parameterize the configuration of interest using a set of design variables.
3. Generate the fine mesh around the modeled configuration.
4. Solve the flow equations for flow variables initiating from upstream flow
variables.
5. Calculate objective functions and constraints.
6. Optimize the current geometry using GA and update design variables.
7. Return to 2. and repeat until an optimum has been reached.
By adding the mesh refinement step to the above algorithm, the design
procedure of the adaptive mesh method can be described as follows
1. Define initial design variables.
2. Parameterize the configuration of interest using a set of design variables.
3. Generate the coarse mesh around the modeled configuration.
4. Solve the flow equations for flow variables initiating from upstream flow
variables.
5. Refine the coarse mesh by evaluating the error indicator from the obtained
flow variables.
6. Interpolate the solution from the coarse mesh to the adapted one.
7. Resolving the flow equations initiating from interpolated flow variables.
8. Calculate objective functions and constraints.
9. Optimize the current geometry using GA and update design variables.
10. Return to 2. and repeat until an optimum has been reached.
A summary of the design process for the both uniform mesh and adaptive
mesh methods is illustrated in Figure 3.1.
3.2
3.2.1
Flow Analysis
CFD Solver Enhanced by Adaptive Refinement
The investigation carried out in this chapter makes use of the PUMI code developed in CIMNE [30]. The reason for this choice is that the focus of this research
work is on the optimization algorithm rather than on the flow solver. Hence, it
is decided to use a validated solver for the flow equations. PUMI is a 3-D flow
solver for high speed inviscid applications. The Galerkin finite element method
in conjunction with the upwind stabilization technique is implemented in this
code. To achieve optimum performance and reduce the memory requirements
an edge based data structure is selected. For time integration, an explicit multistage Runge-Kutta scheme is chosen in order to increase the allowable time step.
More information about the PUMI CFD solver can be found in [30].
70
Define Initial
Design Variables
Parameterize
Configuration
Uniform Mesh
Adaptive mesh
Mesh
Generate The Coarse
Mesh
Generate The Fine
Mesh
Solve Flow
Equations
Mesh Refinement
(Remeshing, interpolating
& resolving)
Solve Flow
Equations
Calculate Objective
Functions & Constraints
Optimization
Optimized?
No
New Design
Variables
Yes
Stop
Figure 3.1: Flowchart of the design process.
71
The main goal of this part of the work is to compare the optimized results
coupled with the uniform and adaptive meshes. The adaptive remeshing technique introduced in Section 2.6 is joined to the CFD solver. Two transonic
test cases are presented in the next section to validate the CFD solver and the
adaptive refinement techniques implemented.
3.2.2
Validation of the mesh refinement technique
In this section some numerical examples are presented in order to illustrate the
performance of the refinement procedures (adaptive remeshing methodologies)
in conjunction with the PUMI solver. The mesh generation is carried out using
the GiD pre/post processing system via an advanced front technique.
All the examples are related to the solution of an inviscid transonic flow over
a NACA 0012 airfoil at different Mach numbers and angles of attack. When
the simulation starts, some time steps are performed in order to get a value of
the density temporal residual of 1.00E − 6. Then, the first refinement level is
done. Consecutive refinement levels are carried out every 200 time steps. The
computational domain is discretized by an unstructured distribution of 2084
nodes and 3970 3-noded triangle elements. The initial mesh is shown in Figure
3.2 which is the same for all the test cases.
Figure 3.2: Initial mesh containing 2084 nodes and 3970 3-noded triangle elements.
Test 1
The first example concerns a subsonic flow with a freestream Mach number
M∞ = 0.8 and angle of attack α = 0.0◦ . After 2, 4, 6, 10, 20 and 30 levels of
remeshing, the final meshes are shown in Figure 3.3. Note that the adaptive
procedure captures all the flow features with precision. The shock wave on the
upper and lower sides of the airfoil and the leading and trailing edge regions
are appropriately captured via the refinement procedure. In Figure 3.4, the Cp
field around the airfoil is introduced using the final adaptive mesh.
72
(a)
(b)
(c)
(d)
(e)
(f)
Figure 3.3: The obtained mesh after (a) 2 (b) 4 (c) 6 (d) 10 (e) 20 (f) 30
remeshing levels for M∞ = 0.8 and α = 0.0◦ .
73
Figure 3.4: Cp field after 20 refinement levels for M∞ = 0.8 and α = 0.0◦ .
Figure 3.5 shows the Cp distribution around the airfoil after 30 levels of
remeshing compared with numerical results [45] where a good agreement is obtained. As shown in Figure 3.6, the variation of total number of nodes and
elements decrease while the refinement goes forward. Assuming 0.001 as the
minimum elem et size, after 30 levels of remeshing, the mesh contains 7782
nodes and 15289 elements. It is noticeable that the total number of nodes and
elements do not change much while the refinement progresses.
Figure 3.5: Comparison between the Cp distribution after 30 remeshing levels
and numerical results reported in [45] for M∞ = 0.8 and α = 0.0◦ .
74
Nodes
Elements
Figure 3.6: Total number of nodes and elements versus the refinement level
number for M∞ = 0.8 and α = 0.0◦ .
Test 2
In this example the freestream Mach number and the angle of attack are considered M∞ = 0.85 and α = 1.0◦ . The refined meshes after 2, 4, 6, 10, 20 and
30 levels of remeshing are shown in Figure 3.7. It is clear that the upper and
lower shockwaves are captured exactly as well as the leading and trailing edges.
Figures 3.8 and 3.9 show the Cp field of the refined mesh and the comparison between obtained Cp distribution with numerical results reported in [49],
respectively. A good agreement is found between the adaptive remeshing results
and the reference ones.
75
(a)
(b)
(c)
(d)
(e)
(f)
Figure 3.7: The obtained mesh after (a) 2 (b) 4 (c) 6 (d) 10 (e) 20 (f) 30
remeshing levels for M∞ = 0.85 and α = 1.0◦ .
Figure 3.8: Cp field after 20 refinement levels for M∞ = 0.85 and α = 1.0◦ .
76
Figure 3.9: Comparison between the Cp distribution after 30 refinement levels
and numerical results reported in [49] for M∞ = 0.85 and α = 1.0◦ .
As shown in Figure 3.10, a suitable convergence is obtained for the total
number of nodes and elements during the remeshing process.
Nodes
Elements
Figure 3.10: Total number of nodes and elements versus the refinement level
number for M∞ = 0.85 and α = 1.0◦ .
77
3.3
Optimization Methodology
Gradient-based methods are well-known optimization algorithms, which find
the optimal design using local gradient information. On the other hand, Genetic Algorithms (GAs) are known to be robust optimization algorithms. Even
though GAs take more computational time to converge to optimal design when
compared to the traditional Gradient-based methods, GAs have capabilities to
escape from local optima and to find the global solution. In addition, GAs are
based on a fitness function evaluations and no gradients are needed during the
optimization process.
Indeed, most of the real-world optimization problems are multi objective
optimization and GAs have unique benefits for solving these type of problems
because multiple Pareto solutions can be obtained simultaneously without specifying weights between objectives [88]. Beside its applications in a wide range
of engineering design problems, GAs have been successfully applied to solve
aerodynamic shape optimization problems [98, 118].
3.3.1
Multi-Objective Genetic Algorithm
In the Origin of Species [22], Charles Darwin stated the theory of natural evolution. Over many generations, biological organisms evolve according to the
principles of natural selection like survival of the fittest to reach some remarkable forms of accomplishment. In nature, individuals in a population have to
compete with each other for vital resources. Because of such selection, poorly
performing individuals have less chance to survive, and the most adapted, or
fit individuals produce a relatively larger number of offsprings. After a few
generations, species evolve spontaneously to become more and more adapted
to their environment. In 1975, Holland developed this idea in Adaptation in
Natural and Artificial system [39]. By describing how to apply the principles
of natural evolution to optimization problems, he laid down the first Genetic
Algorithm. Holland0 s theory has been further developed and now Genetic Algorithms (GAs) accepted as a powerful adaptive method to solve search and
optimization problems. Some evidences of these results can be found in [72, 71].
In GAs, we use the term individual to denote one configuration of the optimal shape. The feasibility of the shape is judged by a fitness function, reflecting
the minimized cost functional and penalizing geometrically unfeasible individuals. The shape parameters of one individual are encoded in the individual0 s
chromosome. The GA operates simultaneously on an entire population of individuals (shapes), the initial population being generated either randomly or as a
set of feasible individuals using an a-priori engineering guess.
GAs are stochastic iterative processes that are not guaranteed to converge.
They have a great potential to explore the whole search space and identify
multiple local maxima and to converge to the global optimum while a point to
point method will generally stall in a local optimum only. They are found more
robust in the case of non differentiable, multi-modal or non convex functions,
and are particularly interesting for search of a trade-off optimum with respect
78
to several criteria (Pareto front, Nash game).
The core of the GA consists in selection, crossover and mutation operators,
whose role is to mimic natural empiric laws of survival of the fittest (selection),
their precreations (crossover) and occasional mutations. The generation operators are usually not deterministic, they implement their operators in the probabilistic sense, with a given probability distribution. The crossover is performed
with a crossover probability pcross (or crossover rate); two selected individuals
(parents) exchange parts of their chromosome to create two offsprings. This
operator tends to enable the evolutionary process to move towards promising
regions of the search space. The mutation operator is introduced to prevent
premature convergence to local optima by randomly sampling new points in the
search space. It is carried out by flipping bits at random, with some (small)
probability pmut .
For multi-objective optimization problems, it is necessary to make some
modifications to the basic GAs. There exist several variants of GAs for this
kind of problem. For the multi-objective GA, the CIMNE in-house software
named Robust Multi-objective Optimization Platform (RMOP) is utilized. It is
a distributed/parallel computational intelligence framework which is a collection
of population based algorithms including GA [24, 53] and Particle Swarm Optimization (PSO) as shown in Figure 3.11. In this work, a GA searching method
in RMOP is used and it is denoted as MOGA. MOGA uses a Pareto tournament
selection operator which ensures that the new individual is not dominated by
any other solutions in the tournament.
Figure 3.11: Topology of Robust Multi-objective Optimization Platform.
For the constraint handling, the linear penalty method is used in such a way
that a weighted sum of the individual constraint violation is added to its fitness
79
value if the constraint is not satisfied. In this work, each generation consists
of 20 individuals and the termination criterion is predefined by the number of
generation. In all the test cases of this research, the initial population is created
randomly.
RMOP is easily coupled to any analysis tools such as Computation Fluid
Dynamic (CFD), Finite Element Analysis (FEA) and/or Computer Aided Design (CAD) systems. Details of RMOP and its engineering design applications
can be found in References [19, 21, 20]. In this work, MOGA is coupled with
both conventional mesh technique and an advanced adaptive mesh refinement
strategy mentioned in Section 3.2.1.
3.3.2
Parametrization
Bezier curves [27] are used to represent the geometry of the airfoil as a linear
combination of the so called Bezier polynomials. Given a set of control points,
the corresponding Bezier curve is defined as


N
X
Xt =
Bi,N (t)Ri
Bi,N =  N  ti (1 − t)N −i
(3.1)
i=0
i
where t ∈ [0, 1] denotes the curve parameter, Bi,N (t) are the Bezier polynomials
of order N and are the coordinates of the control points. The different smooth
curves are created by changing these control points.
The geometries of the airfoils used in this work (NACA 0012 and RAE 2822)
are represented using 24 control points as shown in Figure 3.12. To represent the
airfoil geometry accurately, a bigger density of control points is placed close to
the zone where the airfoil curve has a bigger curvature. The y coordinates of the
control points are considered as the design variables while fixing x coordinates.
Two control points, at the leading edge and trailing edge, have fixed values
during the optimization to keep the chord length constant. Also, two other
control points near the leading edge are fixed to obtain enough curvature in
that zone. In total, 20 design variables are considered for the optimization test
cases.
Figure 3.12 also shows the upper and lower bounds for the design variables
corresponding to both airfoils. This defines a wide range of different geometries
which are sufficient for the optimization.
80
Figure 3.12: Upper and lower bound of design variables for NACA 0012 (top)
and RAE 2822 (bottom) in comparison to the original ones and corresponding
free and fixed control points.
3.4
Realistic Optimization Test Cases
Having introduced the components of the current optimization algorithm in
Section 3.2 and Section 3.3, the obtained results are presented in this section
in order to demonstrate the application of mesh refinement in transonic airfoil
design optimization for reconstruction design and drag minimization and multiobjective design. Adaptive remeshing, described in Section 2.6, is selected as
the refinement methodology in all the results presented in this section.
In this research the flight conditions of Mach number and angle of attack
are treated as constant values with a thickness constraint. To demonstrate
the benefit of mesh refinement, the results obtained with and without adaptive
remeshing are compared. By assuming fixed values for the minimum element size
and the number of refinement steps in the process of adaptive remeshing, it is
possible to limit the computational cost of adaptive remeshing. By investigating
the computational costs related to some random geometries in the assumed flight
conditions, it can be found that the computational time do not vary more than
2% around the average one.
Hence, the uniform mesh is constructed in such a way that the same compu81
tational cost as the average value of the adaptive mesh tests would be required.
Therefore, the characteristics of uniform mesh (such as minimum element size
and the variation of element size far from the airfoil) are defined a priori and
they are considered the same for all the individuals of each optimization problem. Hence, the computational costs for the CFD analysis of the uniform and
adaptive meshes are the same in both reconstruction and drag minimization
design problems. It is notable that the additional costs due to the adaptive
algorithm and the grid regeneration algorithm are also taken into account.
A very fine uniform mesh is considered as the baseline mesh to compare
the final results presented in this research for both reconstruction and drag
minimization designs. A mesh independent study has been carried out in several
flight conditions in order to find a suitable baseline mesh. The minimum element
size considered for the baseline mesh is 2.5 times finer than the one implemented
for the uniform mesh and adaptive mesh used during the optimization. Indeed,
it is dense enough in the vicinity of the airfoil to predict the results in an accurate
manner. Figure 3.13(a) exhibits the baseline mesh on NACA 0012 consisting of
23968 nodes and 47006 3-noded triangle elements and that one on RAE 2822 in
shown in Figure 3.13(b) including 24017 nodes and 47091 elements.
(a)
(b)
Figure 3.13: The baseline mesh around NACA 0012 5(a) and RAE 2822 5(b).
3.4.1
Reconstruction Design
Statement of the problem
This test case considers a single-objective reconstruction design of NACA 0012
at the flow conditions M∞ = 0.78 and α = 2.0◦ . The main objective is to
minimize the pressure error between the target pressure coefficient and the candidate one. The fitness function (the least square error of the pressure) is shown
in Equation 3.2.
f = Pressure Error =
82
N
1 X
(Cp − Cp∗ )2
N i=1
(3.2)
where N represents the number of pressure points on the airfoil (N = 200). As
shown in section 2.3.2, the y coordinate of the control points assumed around
the airfoil are considered as the design variables.
The target pressure coefficient Cp∗ shown in Figure 3.14 is obtained by constructing the baseline mesh around NACA 0012. In the assumed flight conditions, a strong shock wave appears at the upper edge of NACA 0012 profile
which makes this test case suitable to investigate the effect of an adaptive mesh
versus uniform mesh during the reconstruction design of the assumed airfoil.
Figure 3.14: contours around NACA 0012 at the flow conditions M∞ = 0.78
and α = 2.0◦ .
Numerical results
Figure 3.15 compares the convergence history for the pressure error obtained
by MOGA coupled with the uniform mesh and adaptive remeshing approaches.
Both test cases were allowed to run for 133 hours and 400 generation using a
single 1 × 2.8 GHz processor. The uniform mesh approach achieves an optimal
airfoil with pressure error of 0.0273 after 222 generations (74 hours). The adaptive remeshing technique coupled with MOGA converged to the pressure error of
0.0213 after 208 generations. This reflects that the adaptive remeshing technique
produces a 22% more accurate solution. In addition, the adaptive remeshing
technique captures the converged value of the uniform mesh method after 99
generations (33 hours) which is only 45% of uniform mesh computational cost.
In other words, the adaptive remeshing method saves 55% of the computational
cost of the uniform mesh approach. The main reason that the adaptive remeshing method can produce accurate solution within low computational cost, is that
the final adapted mesh conditions provide a better environment to simulate flow
phenomena when compared to the uniform mesh approach.
83
Figure 3.15: Convergence history of reconstruction design of uniform and adaptive mesh test cases.
Table 3.1 compares the final pressure errors obtained for the uniform and
the adaptive mesh test cases. It is shown that the adaptive remeshing test
case airfoil produces 22% lower pressure error than the one resulting from the
uniform mesh test case.
Table 3.1: Comparison of fitness function values obtaining for adaptive mesh
and uniform mesh test cases.
Pressure error
Adaptive mesh
0.0213 (-22%)
Uniform mesh
0.0273
It is noticeable that the pressure errors presented in Figure 3.15 are obtained
by constructing the uniform mesh and the adapted mesh on the optimized airfoils. Since these two meshes are completely different, it can be found that a
portion of the difference in the final pressure error is due to the use of different
meshes.
In order to neglect the effect of mesh difference on the comparison of the
computational cost, the baseline mesh is utilized. By constructing the baseline
mesh on the final airfoil obtained from uniform mesh test case, the pressure error
of 0.023 is obtained. It can be found that this value is captured by constructing
84
the baseline mesh on the airfoil obtained from the adaptive mesh test case after
144 generations which is 65% of the computational cost of the uniform mesh
test case. This expresses that the adaptive remeshing technique improves the
efficiency of the optimization by 35%.
Also, In order to have a fairer comparison on the pressure error, the baseline
mesh is constructed on each optimized airfoil to obtain the final pressure error.
The corresponding values using the baseline mesh for both optimized airfoils
are presented in Table 3.2 which shows a reduction of the order of 24% in the
pressure error.
Table 3.2: Comparison of fitness function values obtaining for adaptive mesh
and uniform mesh cases eliminating the effect of mesh difference.
Pressure error
Adaptive mesh
0.0174 (- 24%)
Uniform mesh
0.0230
The geometries of the target and optimal airfoils obtained by the uniform
mesh and adaptive remeshing approaches are compared in Figure 3.16. Even
though both approaches can capture the target geometry, the adaptive remeshing technique produces a geometry which has a good agreement to the target
airfoil when compared to the optimal airfoil using the uniform mesh approach.
Figure 3.16: Comparison between target, uniform mesh and adaptive mesh test
case airfoils.
Cp distributions corresponding to target airfoil and two optimized test cases
by constructing the baseline mesh for the discretization of each domain are
shown in Figure 3.17. It can be seen that the target pressure distribution on
the lower surface of the airfoil is achieved by the uniform and adaptive mesh
test cases while the adaptive mesh one creates a closer pressure distribution to
the target one on the upper surface due to the more accurate results obtained
85
from adaptive mesh during the optimization. It is noticeable that the shock
wave is captured by both methodologies in an accurate manner.
Figure 3.17: Comparison between target, uniform mesh test case and adaptive
mesh test case Cp distributions.
The corresponding Cp contours and the Mach number contours for the target
airfoil and the two optimized test cases are presented in Figures 3.18 and 3.19.
A good agreement in the pressure and Mach numbers and their corresponding
minimum/maximum values around the airfoils is obtained versus the target one.
86
(a)
(a)
(b)
(b)
(c)
(c)
Figure 3.18: Cp contours around target
airfoil (a), uniform mesh test case (b)
and adaptive mesh test case (c) in Cp
range of [−1.21 : 1.19].
87
Figure 3.19: Mach number contours
around target airfoil (a), uniform mesh
test case (b) and adaptive mesh test
case (c) in Mach range of [0.0 : 1.38].
3.4.2
Drag Minimization
Statement of the problem
The test case considers a single-objective design problem for RAE 2822 to
improve aerodynamic efficiency at the fixed flow conditions M∞ = 0.75 and
α = 3.0◦ . This flow conditions have been selected in order to avoid a possible
shock free solution of the optimization problem. The presence of the shock wave
justifies the use of the technique proposed in this work. The fitness function is
shown in Equation 3.3 with thickness constraint 3.4.
f=
1
L/D
(3.3)
subject to
t
( )max ≥ 0.12
c
(3.4)
where L/D and ct represent the lift to drag ratio and the thickness ratio of the
airfoil, respectively.
In other words, the inverse of lift to drag ratio is minimized at specified
flight conditions while maintaining the maximum thickness of the baseline design (RAE 2822). The design variables selected for drag minimization are the
same as the ones assumed in the reconstruction design.
Numerical results
As illustrated in Figure 3.20, both MOGA coupled with the uniform mesh and
the adaptive remeshing techniques are allowed to run for 150 generations (50
hours) using a single 4 × 2.8 GHz processor. The uniform mesh test case has
converged to f = 0.02979 after 142 generations (47.3 hours). This value is
captured by the adaptive remeshing approach after 8 generations (2.67 hours).
In other words, the adaptive remeshing method improves the optimization efficiency by 95% when compared to the adaptive remeshing technique.
88
Figure 3.20: Convergence history for drag minimization for the uniform and
adaptive mesh test cases.
The fitness value resulted from the target design and the optimized ones
are compared in Table 3.3. It is clear that both optimal airfoils obtained by
the uniform and adaptive mesh techniques improve the aerodynamic behavior
significantly. By comparing the final fitness function of uniform and adaptive
test cases, it can be found that the adaptive one improves the aerodynamic
performance by 48% when compared to the baseline design while the uniform
one only can make improvements of the order of 30%. This improvement on the
reduction of objective function is because of the better CFD results obtained
from the adaptive mesh during the optimization process.
Table 3.3: Comparison of fitness function values obtained from baseline RAE
2822, adaptive mesh and uniform mesh test cases.
1
L/D
Baseline
0.04232
Adaptive mesh
0.02221 (-48%)
Uniform mesh
0.02979 (-30%)
As the reconstruction test case, in order to eliminate the effect of the mesh
89
quality on the values presented in Figure 3.20 and Table 3.3, the objective functions can be recalculated by constructing the baseline mesh on each optimized
airfoil. This leads to a fairer comparison on the final objective function for the
uniform mesh and the adaptive mesh test cases.
The final objective function corresponding to the final airfoil resulted from
uniform mesh test case has the order of 0.0242 if the baseline mesh is used for
the discretization of the domain. This value is captured by constructing the
baseline mesh on the optimized airfoil obtained from the adaptive mesh test
case on the 36th generation. This computational cost is only 25% of the one for
the uniform mesh test case. This shows that the adaptive remeshing technique
improves the efficiency of the optimization by 75%.
The objective function corresponding to the final airfoils resulting from both
test cases (using the baseline mesh) are presented in Table 3.4. It can be seen
that the adaptive remeshing increases the quality of the optimal airfoil by 53%
and 20% when compared to the baseline design and the optimal airfoil obtained
by using the uniform mesh. Hence, even though the baseline mesh is used for
each optimized airfoil, the adaptive mesh test case airfoil yields a more improved
objective function.
Table 3.4: Comparison of fitness function values obtained from baseline RAE
2822, adaptive mesh and uniform mesh cases eliminating the effect of mesh
difference.
1
L/D
Baseline
0.0439
Adaptive mesh
0.02050 (-53%)
Uniform mesh
0.0242 (-45%)
The geometries of baseline and optimal airfoils obtained by the uniform and
adaptive mesh techniques are compared as shown in Figure 3.21. The difference
between the baseline airfoil and the optimized ones can be seen in this figure.
The effect of adaptive remeshing on the optimized airfoil is different in the lower
surface as well as in the upper surface.
90
Figure 3.21: The comparison between the baseline RAE 2822, the adaptive
mesh and uniform mesh test case.
Table 3.5 compares the airfoil characteristics such as the maximum thickness,
maximum camber for the baseline design and optimal airfoils for the uniform and
adaptive mesh techniques. Both optimal airfoils obtained using the uniform and
adaptive mesh techniques have lower camber while maintaining similar thickness
ratio when compared to the baseline design.
Table 3.5: Airfoil configuration of the baseline RAE 2822 and optimized airfoils
using adaptive mesh and uniform mesh.
(t/c)max
(camber)max
Baseline
12.11% (@ 37%)
1.26% (@ 75%)
Adaptive mesh
12.14% (@ 37%)
1.17% (@ 78%)
Uniform mesh
12.48% (@ 37%)
1.11% (@ 26%)
Cp distributions obtained with the baseline design, uniform mesh and adaptive mesh techniques are compared in Figure 3.22 using the baseline mesh for
each one. It can be seen that both test cases reduce the intensity of the shock
wave efficiently while the optimal airfoil obtained with the adaptive remeshing technique has a weaker shock wave, when compared to the uniform mesh
optimal solution.
91
Figure 3.22: Cp distribution obtained from baseline, adaptive mesh and uniform
mesh test cases.
Figures 3.23 and 3.24 compare Cp and Mach number contours obtained by
the baseline design and the optimal airfoils from the optimization. It can be seen
that the strong shock wave on the upper surface of the baseline design is weaker
in the optimized airfoil geometry especially in the lower camber (as deduced
from Table 3.5). Therefore, the drag minimization approach has succeeded to
decrease the strength of the shock wave on the upper surface of the baseline
design.
92
(a)
(a)
(b)
(b)
(c)
(c)
Figure 3.23: Cp contours around original airfoil (a), uniform mesh test case
(b) and adaptive mesh test case (c) in
Cp range of [−1.45 : 1.17].
93
Figure 3.24: Mach number contours
around original airfoil (a), uniform mesh
test case (b) and adaptive mesh test
case (c) in Mach range of [0.0 : 1.43].
3.4.3
Multi-Objective Design
Statement of the problem
In this case, a multi-objective transonic airfoil shape design optimization is
conducted to improve the transonic aerodynamic characteristics of a NACA
0012 profile especially the lift to drag ratio (L/D) and lift coefficient (Cl).
The fitness functions are defined as shown in Equations 3.5 where L/D and Cl
are maximized to extend the aircraft range and to improve its maneuverability,
respectively. The optimization has two constraints for geometry (thickness ratio:
t/c) and aerodynamic performance (Clmin ). The geometry constraint is to
maintain the fuel tank size of the aircraft and the aerodynamic performance
constraint is to keep a flight level at flight conditions of M∞ = 0.75 and α = 3.0◦ .
This example involves the minimization of two objective functions
1
L/D
1
f2 =
Cl
(3.5)
t
( )max ≥ 0.12
c
Cl ≥ 0.45819
(3.6)
f1 =
subject to
The lift coefficient constant is calculated using Equation 3.7 which represents
the minimum lift coefficient for the aircraft at level flight, as
Cl∞ =
2W
ρV 2 S
(3.7)
where W is the weight force (m × g) of the aircraft: mass m = 77, 564kg
and acceleration of gravity g = 9.81m/s2 , ρ is the air density at 35, 000f t:
ρ = 0.41kg/s3 , S is the wing area: S = 124.58m2 . In this case, the design
parameters for the NACA0012 airfoil defined in Section 3.3.2 are used. Twenty
design variables for the airfoil design are considered in total.
Numerical results
Two optimization algorithms using the MOGA technique coupled with uniform and adaptive remeshing procedures have run over 50 hours of computer
time (150 generations). Figure 3.25 compares the Pareto front obtained by the
baseline design (NACA0012 airfoil), uniform mesh and adaptive remeshing approaches. It can be seen that all Pareto members obtained by both methods
dominate the baseline design and the Pareto front obtained by MOGA technique
with adaptive remeshing technique has a better convergence when compared to
the Pareto front obtained by MOGA coupled with uniform mesh. Also, the
use of adaptive remeshing technique speeds up the optimization process while
improving the solution accuracy. From the Pareto front obtained by MOGA
94
with adaptive remeshing techniques, the best solutions (Pareto members 1 and
40) for fitness functions 1 and 2, and one of the compromised solutions (Pareto
member 25) are selected for further comparisons.
Figure 3.25: Optimized Pareto fronts after 150 generations and baseline design.
Table 3.6 compares the aerodynamic characteristics obtained by the baseline
design, Pareto members 1, 25 and 40. The best solution for the fitness function
1 (Pareto member 1) reduces the total drag (Cd) by 51.97% while improving
112% the lift to drag ratio. The best solution for the fitness function 2 (Pareto
member 40) produces 34.01% and 21.46% higher Cl and L/D. The compromised
solution (Pareto member 25) produces 23.69% higher Cl and 23.29% lower Cd
which results in 61.21% L/D improvement.
Table 3.6: The comparison of the aerodynamic coefficients for adaptive remeshing approach and the baseline design NACA 0012.
Baseline
Pareto M1
Pareto M25
Pareto M40
Cl
0.5457
0.5540 (+1.52%)
0.6750 (+23.69%)
0.7313 (+34.01%)
Cd
0.0279
0.0134 (-51.97%)
0.0214 (-23.29%)
0.0341 (+22.22%)
95
L/D
19.57
41.49 (+112.00%)
31.55 ( +61.21%)
21.46 ( +9.66%)
Figure 3.26 compares the geometries obtained by the baseline design and
Pareto members 1 (the best solution for f1), 25 (compromised solution) and 40
(the best solution for f2). It can be seen that Pareto optimal solutions 1, 25 and
40 have slightly higher camber compared to the baseline design. The geometric characteristics; maximum thickness ration (t/c) and maximum camber, are
compared in Table 3.7. It can be found that there is similarity on the thickness
ratio and its position due to the first geometry constraint, while three solutions
have different maximum cambers and its positions.
Figure 3.26: Geometry comparison between the baseline NACA 0012 and optimal airfoils.
Table 3.7: Airfoil configuration of the baseline NACA 0012 and the optimized
airfoils using adaptive remeshing approach.
(t/c)max
(camber)max
Baseline
12.00% (@30%)
0
Pareto M1
12.03% (@35%)
0.86% (@15%)
Pareto M25
12.01% (@ 33%)
0.78% (@67.44%)
Pareto M40
12.08% (@32%)
1.11% (@56%)
Figure 3.27 compares the pressure coefficient (Cp ) distributions obtained by
the baseline design and the compromised solution (Pareto member 25). The
same comparison for the Cp contours has been performed in Figures 3.28a and
3.28b. It can be seen that the shock position on the suction side of airfoil is
moved towards the trailing edge when compared to the baseline design while
reducing the strength of the shock. It results in improving the values of Cl and
L/D by 23.69% and 61.21%, respectively.
96
Figure 3.27: Cp distribution obtained from baseline design and compromised
solution.
(a)
(b)
Figure 3.28: Cp contours obtained by the baseline design (a) and compromised
solution (b) at Cp range of [−1.37 : 1.20].
97
Chapter 4
Concluding Remarks and
Future Work
In this chapter, the conclusions from the current work in this work and some
recommendations for future work will be presented. They can be divided into
two categories: contributions in terms of the stabilized FE formulation aspects
and contributions in terms of the aerodynamic shape optimization aspects.
Part I: Development of a Stabilized Formulation For Compressible
Flows
Based on the concept of Finite Calculus (FIC), a new algorithm for stabilization of the compressible Euler and Navier-Stokes equations is presented in
two-dimensional domains for linear structured/unstructured triangles. Different types of numerical examples have been studied to validate the features of
the new methodology for subsonic, transonic and supersonic flows. Some results have been compared to the analytical values, whereas others have been
compared to some reference numerical solutions.
The developed FIC-FEM stabilized formulation has led to stable and accurate solutions in regions where the flow has some complexities such as shock
wave, boundary layer, stagnation point, etc. It was found that shocks were
resolved within four or five elements. Indeed, the boundary layer is captured
through an accurate velocity profiles in the boundary layer zone as well as the
appropriate pressure coefficient Cp distribution and the skin-friction coefficient
Cf distribution along the boundary.
For inviscid flows, error estimation and adaptive mesh refinement have improved the relation cost and accuracy, especially for the flows with shocks. In
the case of viscous flows, the formulation has worked well using structured elements with maximum aspect ratio about 10, except in the zone of singularity
where the formulation is not able to control the temperature causing a increase
in the Mach number, locally.
For future work, we intend to enhance the accuracy of the proposed formula-
98
tion for estimating the temperature inside the elements with high aspect ratio.
Consideration of the turbulence inside the governing equations can advance the
methodology for solving more realistic flow problems. The extension of the current formulation to three-dimensional domains is another target of the future
developments in the research.
Part II: Aerodynamic Shape Optimization Using Adaptive Remeshing
In the second part of the work, a methodology coupling the MOGA and an
adaptive remeshing technique has been developed and has been implemented
for three practical aerodynamic shape optimization problems in a single- and
multi-objective manner. From numerical studies, both the MOGA technique
coupled to uniform mesh and adaptive remeshing approaches have been validated through reconstruction/inverse test case and implemented to solve two
complex aerodynamic shape design optimizations, namely, drag minimization
and multi-objective shape design. It is noticed that despite using the same
MOGA technique and CFD analyzer, the use of adaptive remeshing approach
accelerates the optimization process while increasing the solution accuracy when
compared to the solution using a uniform mesh technique.
Future works will focus on implementation of the adaptive remeshing technique in a multi-point design optimization considering uncertain design parameters (robust/uncertainty). Also, by implementing gradient-based optimization
methodologies, the capability of the adjoint method for predicting sensitivities
of the objective function can be studied. Indeed, the development of the current
method for three-dimensional optimization test cases will improve the generalization of this optimization framework.
99
Bibliography
[1] M. Aftosmis and N. Kroll. A quadrilateral based second-order tvd method
for unstructured adaptive meshes. AIAA-91-0124, 1991.
[2] W.K. Anderson, J.L. Thomas, and B. Van Leer. Comparison of finite
volume flux vector splittings for the euler equations. AIAA Journal, 24
(9):1453–1460, 1986.
[3] T. J. Baker. Developments and trends in three-dimensional mesh generation. Applied Numerical Mathematics, 5:275–304, 1989.
[4] J. P. Boris and D. L. Book. Flux corrected transport i. shasta, a fluid
transport algorithm that works. J. Comput. Phys., 11:38–69, 1973.
[5] A.N. Brooks and T.J.R. Hughes. Streamline upwind/petrov-galerkin formulations for convection dominated flows with particular emphasis on the
incompressible navier-stokes equations. Comput. Methods Appl. Mech.
Engrg, 32:199–259, 1982.
[6] G. Bugeda and E. Oñate. Optimum aerodynamic shape design including
mesh adaptivity. International Journal for Numerical Methods in Fluids,
20:915–934, 1995.
[7] G. Bugeda and E. Oñate. Optimum aerodynamic shape design for fluid
flow problems including mesh adaptivity. International Journal for Numerical Methods in Fluids, 30:161–178, 1998.
[8] N. K. Burgess and D. J. Mavriplis. An hp-adaptive discontinuous galerkin
solver for aerodynamic flows on mixed-element meshes. In Proceeding of
the 49th Aerospace Sciences Meeting, Orlando, FL, Jan 2011. AIAA Paper
2011-490, 2011.
[9] G. W. Burgreen and O. Baysal. Three-dimensional aerodynamic shape
optimization of wings using sensitivity analysis. In 32th AerospacePierce
Sciences Meeting and Exhibit, RENO, NEVADA, 1994.
[10] H. V. Cao and G. A. Blom. Navier-stokes/genetic optimization of multielement airfoils. In 14th AIAA Applied Aerodynamics Conference, New
Orleans, LA, 1996.
100
[11] J.E. Carter. Numerical solutions of the navier-stokes equations for the
supersonic laminar flow over a two-dimensional compression corner. NASA
Technical Report, TR R-385, 1972.
[12] F. Chalot and T.J.R. Hughes. A consistent equilibrium chemistry algorithm for hypersonic flows. Comput. Methods Appl. Mech. Engrg, 112:25–
40, 1994.
[13] A. J. Chorin. Numerical solution of the navier stokes equations. Math.
Comput., 22:745–762, 1968.
[14] A. J. Chorin. Sur l’approximation de la solution des equations de navierstokes par la methode des pas fractionaires (i). Arch. Rat. Mech. Anal.,
32:135–153, 1969.
[15] R. Codina. A discontinuity-capturing crosswind-dissipation for the finite
element solution of the convection diffusion equation. Comput. Methods
Appl. Mech. Eng, 110:325–342, 1993.
[16] R. Codina, M. Vasquez, and O.C. Zienkiewicz. A general algorithm for
compressible and incompressible flow. part iii: The semi-implicit form.
Int. J. Numer. Methods Fluids., 27:13–32, 1998.
[17] R. Courant and E. Isaacson. On the solution of nonlinear hyperbolic differential equations by finite differences. Comm. Pure and Applied Mathematics, 5:243–255, 1952.
[18] A. Jameson D. J. Mavriplis and L. Martinelli. February 1989.
[19] E. Onate D. S. Lee, J. Periaux and L. F. Gonzalez. Advanced computational intelligence system for inverse aeronautical design optimization.
International Conference on Advanced Software Engineering (ICASE-11),
Proceedings of the 9th IEEE International Symposium on Parallel and Distributed Processing with Applications Workshops, ISPAW 2011 - ICASE
2011, SGH 2011, GSDP 2011, pp. 299-304, ISBN (978-1-4577-0524-3),
DOI (10.1109/ISPAW.2011.46), Busan, Korea, 26-28 May, 2011.
[20] G. Bugeda S. Oller E. Onate D. S. Lee, C. Morillo. Multilayered composite
structure design optimization using distributed/parallel multi-objective
evolutionary algorithms. Journal of Composite Structures, 94(3):1087–
1096, 2012.
[21] J. Periaux D. S. Lee, G. Bugeda and E. Onate. Robust active shock
control bump design optimization using parallel hybrid-moga. The 23rd
International Conference on Parallel Computational Fluid Dynamics 211,
Barcelona, Spain, 16-20 May, 2011.
[22] C. Darwin. On the origin of species by means of natural selection. John
Murray, London, 1859.
101
[23] S.F. Davis. Tvd finite difference schemes and artificial viscosity. ICASE
Report No. 84-20, 1984.
[24] Goldberg D.E. Genetic Algorithm in Search, Optimization and Machine
Learning. Addison-Wesley, Massachusetts, 1989.
[25] J. Donea. A taylor-galerkin method for convective transport problems.
Int. J. Numer. Methods Engrg, 20:101–120, 1984.
[26] B. Engquist and S. Osher. Stable and entropy satisfying approximations
for transonic flow calculations. Mathematics of Computation, 34:45–75,
1980.
[27] I.D. Faux and M.J. Pratt. Computational Geometry for Design and Manufacture. Ellis Horwood Limited, 1887.
[28] I. Fejtek, D. Jones, G. Waller, E. Hansen, and S. Obayashi. A transonic
wing inverse design capability for complete aircraft configurations. In 15th
AIAA Computational Fluid Dynamics Conference, Anaheim, CA,, 2001.
[29] K. J. Fidkowski and D. L. Darmofal. A triangular cut-cell adaptive method
for high-order discretizations of the compressible navier-stokes equations.
Journal of Computational Physics, 225(2):1653–1672, 2007.
[30] R. Flores, E. Ortega, and E. Oate. Pumi: an explicit 3d unstructured
finite element solver for the euler equations. Technical report, Publication
CIMNE, PI 326, 2008.
[31] M. B. Giles and M. Drela. Two-dimensional transonic aerodynamic design
method. AIAA Journal, 25:1199–1206, 1987.
[32] T.E. Tezduyar G.J. Le Beau. Finite element computation of compressible
flows with the supg formulation. in: Advances in Finite Element Analysis
in Fluid Dynamics, ASME, New York, 123:21–27, 1991.
[33] S. K. Godunov. A difference scheme for numerical computation of discontinuous solution of hydrodynamic equations. Matematicheskii Sbornik,
47:271–306, 1959.
[34] A. Harten. High resolution schemes for hyperbolic conservation laws. J.
Comput. Phys., 49:357–393, 1983.
[35] O. Hassan. Finite element computation of high speed viscous compressible
flows. PhD Thesis, University of Wales Swansea, 1990.
[36] G. Hauke and T.J.R. Hughes. A unified approach to compressible and
incompressible flows. Comput. Methods Appl. Mech. Engrg, 113:389–395,
1994.
[37] R. M. Hicks and P. A. Henne. Wing design by numerical optimization.
Journal of Aircraft, 15(7):407–412, 1977.
102
[38] C. Hirsch. Numerical Computation of Internal and External Flow, vol. 1.
J. Wiley, New York, 1988 (vol. 2, 1990).
[39] J. H. Holland. Adaptation in Natural and Artificial Systems: An Introductory Analysis with Applications to Biology, Control, and Artificial
Intelligence. University of Michigan Press, 1975.
[40] T. L. Holst and T. H. Pulliam. Aerodynamic shape optimization using
a real- number-encoded genetic algorithm. In 15th AIAA Computational
Fluid Dynamics Conference, Anaheim, CA,, 2001.
[41] T.J.R. Hughes, L.P. Franca, and M. Mallet. A new finite element formulation for computational fluid dynamics: Vi. convergence analysis of the
generalized supg formulation for linear time-dependent multi-dimensional
advective-diffusive systems. Comput. Methods Appl. Mech. Engrg, 63:97–
112, 1987.
[42] T.J.R. Hughes and T.E. Tezduyar. Finite element methods for first-order
hyperbolic systems with particular emphasis on the compressible euler
equations. Comput. Methods Appl. Mech. Engrg, 45:217–284, 1984.
[43] C.M. Hung and R.W. MacCormack. Numerical solutions of supersonic and
hypersonic laminar compression corner flows. AIAA J., 14 (4):475–481,
1976.
[44] K. Morgan J. Peraire and J. Peiro. Unstructured mesh methods for cfd.
Technical Report 9004, Imperial College Aeronautics Department, 1990.
[45] A. Jameson. The evolution of computational methods in aerodynamics.
Journal of Applied Mathematics, 50:1052–1070, 1983.
[46] A. Jameson. Aerodynamic design via control theory. Journal of Scientific
Computing, 3:233–260, 1988.
[47] A. Jameson. Computational aerodynamics for aircraft design. Science,
245:361–371, 1989.
[48] A. Jameson. Automatic design of transonic airfoils to reduce the shock
induced pressure drag. In Proceedings of the 31st Israel Annual Conference
on Aviation and Aeronautics, Tel Aviv, pages 5–17, 1990.
[49] A. Jameson. Positive schemes and shock modeling for compressible flows.
International Journal for Numerical Methods in Fluids, 20:743–776, 1995.
[50] A. Jameson and P.D. Lax. Conditions for the construction of multi-point
total variation diminishing difference schemes. MAE Report 1650, Dept .
of Mechanical and Aerospace Engineering, Princeton University, 1984.
[51] A. Jameson, L. Martinelli, and N. Pierce. Optimum aerodynamic design
using the navier-stokes equations. In 35th Aerospace Sciences Meeting and
Exhibit, RENO, NEVADA, 1997.
103
[52] A. Jameson, W. Schmidt, and E. Turkel. Numerical solution of the euler equations by finite volume methods using runge-kutta time stepping
schemes. AIAA paper, 81-1259, June 1981.
[53] S. Agrawal K. Deb, A. Pratap and T. Meyarivan. A fast elitist nondominated sorting genetic algorithm for multi-objective optimization:
Nsga-ii. IEEE Transactions on Evolutionary Computation, 6 (2):182197,
2002.
[54] D. Kuzmin and S. Turek. High-resolution fem-tvd schemes based on a
fully multidimensional ux limiter. J. Comput. Phys., 198:131–158, 2004.
[55] P. D. Lax, , and B. Wendroff. Difference schemes for hyperbolic equations with high order of accuracy. Comm. Pure and Applied Mathematics,
17:381–398, 1964.
[56] P.D. Lax. Weak solutions of non-linear hyperbolic equations and their
numerical approximations. Communications on Pure and Applied Mathematics, 7:159–193, 1954.
[57] A. Lerat and R. Peyret. Non centered schemes and shock propagation
problems. Computers and Fluids, 2:35–52, 1969.
[58] R. J. LeVeque. Numerical methods for conservation laws (Second ed.).
Birkhauser, 1992.
[59] R. H. Liebeck. A class of airfoils designed for high lift in incompressible
flow. Journal of Aircraft, 10:610–617, 1973.
[60] M. J. Lighthill. A new method of two-dimensional aerodynamic design.
2112, Aeronautical Research Council, 1945.
[61] R. Lohner. An adaptive finite element solver for transient problems with
moving bodies. Comput. and Structures, 30:303–317, 1988.
[62] R. Lohner. Three-dimensional grid generation by the advancing front
method. International Journal for Numerical Methods in Fluids, 8:1135–
1149, 1988.
[63] R. Lohner. Mesh adaptation in fluid mechanics. Engineering Fracture
Mechanics, 50(5/6):819–847, 1995.
[64] R. Lohner and J. D. Baum. Adaptive h-refinement on 3-d unstructured
grids for transient problems. International Journal for Numerical Methods
in Fluids, 14(12):1407–1419, 1992.
[65] R. Lohner, K. Morgan, J. Peraire, and M. Vahdati. Finite element fluxcorrected transport (fem-fct) for the euler and navier-stokes equations.
Int. J. Num. Methods Fluids, 7:103–109, 1987.
104
[66] H. Lomax, T. H. Pulliam, and D. W. Zingg. Fundamentals of Computational Fluid Dynamics. Springer-Verlag, Berlin, Heidelberg, 2001.
[67] H. Luo, J. D. Baum, R. Lohner, and J. Cabello. Adaptive edge-based finite
element schemes for the euler and navier-stokes equations. AIAA-93-0336,
1993.
[68] P. R. M. Lyra1, K. Morgan1, J. Peraire, and J. Peir. Tvd algorithms for
the solution of the compressible euler equations on unstructured meshes.
Int. J. Numer. Methods Flu., 19(9):827–847, 1994.
[69] R.W. MacCormack. The effect of viscosity in hypervelocity impact cratering. AIAA paper 69-0354, April-May, 1969.
[70] K. Mani and D. J. Mavriplis. unsteady discrete adjoint formulation for
two-dimensional flow problems with deforming meshes. AIAA Journal,
46-6:13511364, 2008.
[71] N. Marco and J.A. Desideri. Numerical solution of optimization test-case
by genetic algorithms. Technical report, INRIA Research Report, No.
3622, 1999.
[72] N. Marco and S. Lanteri. A two-level parallelization strategy for genetic
algorithms applied to shape optimum design. Technical report, INRIA
Research Report, No. 3463, 1998.
[73] D. J. Mavriplis. Adaptive meshing techniques for viscous flow calculations on mixed element unstructured meshes. International Journal for
Numerical Methods in Fluids, 34(2):93–111, 2000.
[74] D.J. Mavriplis. Adaptive mesh generation for viscous flows using delaunay
triangulation. Journal of Computational Physics, 90:271–291, 1990.
[75] D.J. Mavriplis. Turbulent flow calculations using unstructured and adaptive meshes. Int. J. Num. Meth. Fluids, 13:11311152, 2008.
[76] K. Miettinen, M. M. Makela, P. Neittaanmaki, and J. Periaux. Evolutionary Algorithms in Engineering and Computer Science. John Willey and
Sons Ltd, Chichester, U.K., 1999.
[77] S. Mittal and T. Tezduyar. A unified finite element formulation for compressible and incompressible flows using augmented conservation variables. Comput. Methods Appl. Mech. Engrg, 161:229–243, 1998.
[78] K. Morgan, J. Peraire, J. Peiro, and O.C. Zienkiewicz. adaptive remeshing
applied to the solution of a shock interaction problem on a cylindrical
leading edge. in P. Stow (ed.), Computational Methods in Aeronautical
Fluid Dynamics, Clarenden Press, Oxford, pages 327–344, 1990.
[79] Von Neumann, J., and R. D. Richtmyer. A method for the numerical
calculation of hydrodynamic shocks. J. Appl. Phys, 21:232–237, 1950.
105
[80] E. Oñate. Derivation of stabilized equations for advective-diffusive transport and fluid flow problems. Comput. Meth. Appl. Mech. Engng.,
151:233–267, 1998.
[81] E. Oñate. A stabilized finite element method for incompressible viscous
flows using a finite increment calculus formulation. Comp. Meth. Appl.
Mech. Eng, 182:355–370, 2000.
[82] E. Oñate. Possibilities of finite calculus in computational mechanics. Int.
J. Num. Meth. Engng, 60:255–281, 2004.
[83] E. Oñate and J. Garcia. A finite element method for fluid-structure interaction with surface waves using a finite calculus formulation. Comput.
Meth. Appl. Mech. Eng., 191:635–660, 2001.
[84] E. Oñate, J. Garcia, and S. Idelsohn. Computation of the stabilization
parameter for the finite element solution of advective-diffusive problems.
Int. J. Num. Meth. Fluids., 25:1385–1407, 1997.
[85] E. Oñate, J. Garcia, and S. Idelsohn. An alpha-adaptive approach for stabilized finite element solution of advective-diffusive problems with sharp
gradients. New Adv. in Adaptive Comp. Met. in Mech., P. Ladeveze and
J.T. Oden (Eds.), Elsevier, 1998.
[86] E. Oñate and M. Manzan. A general procedure for deriving stabilized
space-time finite element methods for advective-diffusive problems. Int.
J. Num. Meth. Fluids., 31:203–221, 1999.
[87] E. Oñate and M. Manzan. Stabilization techniques for finite element analysis of convection diffusion problems. in Computational Analysis of Heat
Transfer, G. Comini and B. Sunden (Eds.), WIT Press, Southampton,
2000.
[88] S. Obayashi and A. Oyama. Three- dimensional aerodynamic optimization with genetic algorithms. In Proceedings of the Third ECCOMAS
Computational Fluid Dynamics Conference, pages 420–424, 1996.
[89] S. Osher. Shock modelling in aeronautics. In, K. W. Morton and M. J.
Baines(eds), NumericalMethodsfor Fluid Dynamics, London: AcademicPress, pages 179–218, 1982.
[90] A. Oyama, S. Obayashi, K. Nakahashi, and T. Nakamura. Euler/navierstokes optimization of supersonic wing design based on evolutionary algorithm. AIAA Journal, 37(10):1327–1329, 1999.
[91] B. Palmerio and A. Dervieux. Application of a fem moving node adaptive
method to accurate shock capturing. First Int. Conf. on Numerical Grid
Generation in CFD, Landshut, W. Germany, July 1417, 1986.
106
[92] A. K. Parrott and M. K. Christie. Fct applied to the 2d finite element
solution of tracer transport by single phase flow in a porous medium.
Numerical Methods for Fluid Dynamics, Oxford: Oxford University Press,
pages 27–53, 1986.
[93] J. Peraire, K. Morgan, and J. Peiro. Unstructured finite element mesh
generation and adaptive procedures for cfd. Technical report, AGARDCP-464, 1990.
[94] J. Peraire, J. Peiro, L. Formaggia, K. Morgan, and O.C. Zienkiewicz.
Finite element euler computations in three dimensions. Int. J. Numer.
Methods Eng., 26:2135–2159, 1988.
[95] J. Peraire, M. Vahdati, K. Morgan, and O. C. Zienkiewicz. Adaptive
remeshing for compressible flow computations. Journal of Computational
Physics, 72:449–466, 1987.
[96] J. Periaux, D.S. Lee, L.F. Gonzalez, and K. Srinivas. Fast reconstruction
of aerodynamic shapes using evolutionary algorithms and virtual nash
strategies in a cfd design environment. Journal of Computational and
Applied Mathematics, 232:61–67, 2009.
[97] T. H. Pulliam and J. T. Barton. Euler computations of agard working
group 07 airfoil test cases. AIAA Paper 85-0018, AIAA 23rd Aerospace
Sciences Meeting, Nev. 1985.
[98] D. Quagliarella and A. D. Cioppa. Genetic algorithms applied to the
aerodynamic design of transonic airfoils. In AIAA Applied Aerodynamics
Conference, 12th, Colorado Springs, pages 686–693, 1994.
[99] O. Soto R. Lhner and Chi Yang. An adjoint-based design methodology
for cfd optimization problems. AIAA-03-0299, 2003.
[100] J. Reuther and A. Jameson. Aerodynamic shape optimization of wing and
wing body configurations using control theory. In 33rd Aerospace Sciences
Meeting and Exhibit, RENO, NEVADA, 1995.
[101] J. Reuther, A. Jameson, J. J. Alonso, M. J. Rimlinger, and D. Saunders.
Constrained multipoint aerodynamic shape optimization using an adjoint
formulation and parallel computers. In 33rd Aerospace Sciences Meeting
and Exhibit, RENO, NEVADA, 1997.
[102] P. L. Roe. Approximate riemann solvers, parameter vectors and difference
schemes. Journal Computational Physics, 43:357–372, 1981.
[103] P. L. Roe. The use of the riemann problem in finite differenceschemes.
Lecture Notes in Physics, 141:354–359, 1981.
[104] J. V. Rosendale. Floating shock fitting via lagrangian adaptive meshes.
Technical report, ICASE Report, No. 94-89, 1994.
107
[105] R. Sanders. On convergence of monotone finite difference schemes with
variable spatial differencing. Math. Comp., 40:91–106, 1983.
[106] W. K. Schnauer. The principle of difference quotients as a key to the
self-adaptive solution of nonlinear partial differential equations. Comp.
Meth. Appl. Mech. Eng., 28:327359, 1981.
[107] F. Shakib. Finite element analysis of the compressible euler and navierstokes equations. PhD thesis, Department of Mechanical Engineering,
Stanford University, 1988.
[108] F. Shakib, T.J.R. Hughes, and Z. Johan. A new finite element formulation
for computational fluid dynamics: X. the compressible euler and navierstokes equations. Comput. Methods Appl. Mech. Engrg, 89:141–219, 1991.
[109] M. S. Shepard and M. K. Georges. Automatic three-dimensional mesh
generation by the finite octree technique. International Journal for Numerical Methods in Engineering, 32:709–749, 1991.
[110] O. Soto and R. Lohner. Cfd optimization using an incomplete-gradient
adjoint approach. AIAA-00-0666, 2000.
[111] E. Turkel. Improving the accuracy of central difference schemes. ICASE
Report 8853, 1988.
[112] B. Van Leer. Towards the ultimate conservative difference scheme. ii.
monotonicity and conservation combined in a second-order scheme. J.
Comput. Phys., 14(4):361–370, 1974.
[113] B. Van Leer. Towards the ultimate conservative difference scheme iii.
upstream-centered finite-difference schemes for ideal compressible flow. J.
Comput. Phys., 23 (3):263–275, 1977.
[114] D. Vendetti and D. Darmofal. Grid adaptation for functional outputs:
Application to two-dimensional inviscid flows. Journal of Computational
Physics, 176:40–69, 2002.
[115] D. Vendetti and D. Darmofal. Anisotropic grid adaptation for functional
outputs: application to two-dimensional viscous flows. Journal of Computational Physics, 187:22–46, 2003.
[116] G. Volpe and R. E. Melnik. The design of transonic airfoils by a wellposed inverse method. International Journal for Numerical Methods in
Engineering, 22:341–361, 1984.
[117] L. Wang and D. J. Mavriplis. Adjoint based h-p adaptive discontinuous
galerkin methods for aerospace applications. In 47th AIAA Aerospace
Sciences Meeting including The New Horizons Forum and Aerospace Exposition, Orlando, Florida, 2009.
108
[118] K. Yamamoto and O. Inoue. Applications of genetic algorithm to aerodynamic shape optimization. In 12th AIAA Computational Fluid Dynamics
Conference, CP956, San Diego, California, pages 43–51, 1995.
[119] H.C. Yee. Construction of implicit and explicit symmetric t v d schemes
and their applications. J. Comput. Phys., 68:151–179, 1987.
[120] H.C. Yee and P. Kutler. Application of second-order accurate total variation diminishing (tvd) schemes to the euler equations in general geometries. NASA TM-85845, 1983.
[121] M. A. Yerry and M. S. Shepard. Automatic three-dimensional mesh generation by the modified-octree technique. International Journal for Numerical Methods in Engineering, 20:1965–1990, 1984.
[122] S.T. Zalesak. Fully multidimensional flux-corrected transport algorithm
for fluids. J. Comput. Phys., 31:335–362, 1979.
[123] O.C. Zienkiewicz and R. Codina. A general algorithm for compressible
and incompressible flow. part i. the split characteristic based scheme. Int.
J. Numer. Methods Fluids., 20:869–885, 1995.
[124] O.C. Zienkiewicz, B.V.K. Satya Sai, K. Morgan, R. Codina, and
M. Vazquez. A general algorithm for compressible and incompressible
flow. part ii. tests on the explicit form. Int. J. Numer. Methods Fluids.,
20:887–913, 1995.
[125] O.C. Zienkiewicz and J. Wu. A general explicit of semi-explicit algorithm
for compressible and incompressible flows. Int. J. Numer. Methods Eng.,
35:457–479, 1992.
109